Next Article in Journal
Epigenetics, Stem Cells, and Autophagy: Exploring a Path Involving miRNA
Previous Article in Journal
The Methylation Patterns and Transcriptional Responses to Chilling Stress at the Seedling Stage in Rice
Previous Article in Special Issue
Synergistic Neuroprotective Effect of Schisandra chinensis and Ribes fasciculatum on Neuronal Cell Death and Scopolamine-Induced Cognitive Impairment in Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dietary Polyphenols: A Multifactorial Strategy to Target Alzheimer’s Disease

1
School of Science, RMIT University, Bundoora, Victoria 3083, Australia
2
Mushroom Research Centre, University of Malaya, 50603 Kuala Lumpur, Malaysia
3
Department of Anatomy, Faculty of Medicine, University of Malaya, 50603 Kuala Lumpur, Malaysia
4
Department of Pharmaceutical Life Sciences, Faculty of Pharmacy, University of Malaya, 50603 Kuala Lumpur, Malaysia
5
Institute of Biological Sciences, Faculty of Science, University of Malaya, 50603 Kuala Lumpur, Malaysia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(20), 5090; https://doi.org/10.3390/ijms20205090
Submission received: 19 September 2019 / Revised: 11 October 2019 / Accepted: 11 October 2019 / Published: 14 October 2019

Abstract

:
Ageing is an inevitable fundamental process for people and is their greatest risk factor for neurodegenerative disease. The ageing processes bring changes in cells that can drive the organisms to experience loss of nutrient sensing, disrupted cellular functions, increased oxidative stress, loss of cellular homeostasis, genomic instability, accumulation of misfolded protein, impaired cellular defenses and telomere shortening. Perturbation of these vital cellular processes in neuronal cells can lead to life threatening neurological disorders like Alzheimer’s Disease, Parkinson’s Disease, Huntington’s Disease, Lewy body dementia, etc. Alzheimer’s Disease is the most frequent cause of deaths in the elderly population. Various therapeutic molecules have been designed to overcome the social, economic and health care burden caused by Alzheimer’s Disease. Almost all the chemical compounds in clinical practice have been found to treat symptoms only limiting them to palliative care. The reason behind such imperfect drugs may result from the inefficiencies of the current drugs to target the cause of the disease. Here, we review the potential role of antioxidant polyphenolic compounds that could possibly be the most effective preventative strategy against Alzheimer’s Disease.

Graphical Abstract

1. Introduction

Deaths due to Alzheimer’s Disease (AD) and other dementias are a major cause of mortality in the elderly worldwide, and the rate is increasing rapidly with a doubling time of 20 years [1]. AD is an age-related neurodegenerative disease that leads to cognitive impairment and death. Neuronal synapsis disruption, accumulation of amyloid plaques in brain, formation of neurofibrillary tangles in neuronal cells, loss of cellular homeostasis and accumulation of oxidative stress are major hallmarks of the disease [2]. However, mitochondrial dysfunction, loss of protein and lipid homeostasis, alterations in biometal distribution, cellular senescence, loss of nutrient sensing and accumulation of misfolded proteins are also associated with the AD [3]. Despite the efforts of more than three decades of research, the precise cause of AD has not been found. Many hypotheses have been made to address the major molecular events in the neuronal cells with AD (refer to Figure 1) [2]. Polyphenolic compounds have been reported to have multiple effects in cells including inducing antioxidant activity, induction of autophagy, restoration of lipid homeostasis, antiproliferative property, anti-proteinopathies, inhibition of choline esterases, anti-inflammatory activity, metal chelation, clearance of lipofuscin and others (refer to Table 1). This review details how polyphenols exert their neuroprotective role at the cellular level helping to prevent and possibly cure AD.

2. Current Therapeutic Approaches Only Target Symptoms of AD

Consideration for drug design against AD has come from the symptoms. Traditional approaches based on cholinergic dysfunction have been highly utilised for treatment of AD [2]. Current FDA approved drugs include donepezil, rivastigmine, galantamine and memantine of which the first three drugs are acetylcholine esterase inhibitors, while memantine targets the N-methyl-D-aspartic receptor (NMDAR) [92,93]. Damage of cholinergic neuronal cells leading to the reduced levels of acetylcholine, a neurotransmitter involved in cognition and synapsis, has been found to be associated with AD [94]. Restoring the levels of acetylcholine in an AD brain has been considered to be the most viable palliative measure. The inhibition of acetylcholine esterase has shown benefits in restoring cognition making it a primary care strategy [95]. Likewise, memantine is a NMDAR antagonist as it selectively inhibits the interaction of glutamate with NMDAR, balancing the excitation by the neurotransmitter. The drug effect comes through the reduction of ionotropic channels in the membrane restoring the balanced influx of calcium and sodium ions which is highly expressed in an AD brain causing excitotoxicity [92,96]. However, the strategy targeting only these extracellular events may not provide substantial protection, as many intracellular processes are also altered during progression of AD.

3. Therapeutic Strategies Based on Targeting Amyloid β and Tau Proteins

Several studies involving novel strategies to multiple molecular processes, have been considered. The most popular one among the various newer approaches is targeting amyloid β, also referred to as anti-amyloid strategy. Amyloid β comprise short polypeptides, 36–43 amino acid long, produced after pre-processing of amyloid precursor protein (APP) by two different enzymes, namely β-secretase (BACE) and γ-secretase [97,98]. BACE cleaves the APP at a specific site followed by the action of γ-secretase resulting in the formation of peptides of length 36-43 amino acids. The most important polypeptide found in the amyloid plaques of the patient’s brain is Aβ42, which is well-known for its adverse effects in different disease models [99]. Conversely, α-secretase can cleave APP at a site within Aβ, creating shorter fragments also called Aα, which is non-amyloidogenic and protective [100,101]. BACE exists in two isoforms, namely BACE1 and BACE2 [102,103]. BACE1 has been considered an important drug target as it is intimately involved in the formation of Aβ [104]. The BACE1 enzyme has the aspartic catalytic residues located at the interface of the N-terminus and C-terminus forming a dyad, one of which acts as an acid and the other one as a base during the proteolysis [105,106]. The recent developments enlightening BACE1 structure and function provided opportunities for in silico molecular docking studies supporting drug design and discovery [107]. Various molecules have been studied and evaluated for their inhibitory action against BACE1 including macrocyclic lactones, hydroethylenes, aminoethylenes, aminoimidazoles, aminobenzthiazines, spiropiperidines, etc. [108,109,110,111,112,113,114].
Inhibition of γ-secretase activity is also an important approach in the anti-amyloid strategy. Inhibiting activity of γ-secretase will affect the Aβ formation and is expected to halt the amyloidogenic progress and associated toxicity. However, the interference with the γ-secretase activity also affects the notch signaling [115]. Development and cellular growth are associated with notch signaling mechanism, which will also be altered by inhibiting the γ-secretase [116]. Considering these side effects of the γ-secretase inhibitors, different sulfones and sulfonamides that do not affect notch signaling have been evaluated for their activity against γ-secretase [117]. An anti Aβ-aggregation approach has also been studied in the effort to find a chemo preventative for AD. Aβ aggregation occurs by the interaction of molecules of monomeric Aβ which further interact with other monomeric forms to produce aggregates [118]. Oligomeric forms of Aβ42 have been reported to be the most toxic species. Very few compounds have been evaluated for their anti-aggregation properties [119,120,121,122]. Aβ clearance, inducing misfolded protein degradation through induction of autophagy and unfolded protein response, is another strategy that could provide protection [123]. Furthermore, vaccines and antibodies against Aβ were also evaluated for their efficacy against AD [120,124,125]. Early vaccines targeting Aβ caused serious side effects of meningoencephalitis in the trial and antibodies are limited by the blood brain barrier as only 0.1% of the antibodies were found to cross it [1]. While Aβ remains an important target, its clearance may have limited benefits as a cure after the disease onset. Despite the limitations of approaches targeting Aβ, early prevention of Aβ formation and its clearance remains a top priority.
Tau neurofibrillary tangles (NFTs) are another important pathophysiological hallmark in addition to the accumulation of the amyloid plaques in the AD brain [126]. In a normal brain, tau protein plays a critical role in cellular integrity by maintaining the microtubules [127]. Tau normally stays in the membrane of axons in phosphorylated form, as it contains 84 amino acid residues where phosphorylation can occur [128]. Hyperphosphorylation of these tau proteins leads to self-interaction and reduces its tendency to bind with the microtubules causing the formation of the NFTs [120,129]. Formation of NFTs is associated with alteration in neuronal plasticity and synapsis [116]. Hyperphosphorylation of tau has been reported to be the major contributor for activation of the microglial cells and astrocytes [130]. Activation of these immunomodulators downstream leads to the release of nuclear factor kappa B (NFκB) and cytokines [131], which cause the brain inflammation associated with AD. Meanwhile, release of the inflammatory mediators like NFκB and interleukins result in the activation of protein kinases in cell, which reinforces the hyperphosphorlation of the tau [132,133]. Some of the important protein kinases reported to cause hyperphosphorylation of tau include mitogen-activated protein kinase (MAPK), cyclin dependent kinase-5, tau protein kinase-I and glycogen synthase kinase-3β (GSK-3β) [120]. Inhibition of these protein kinases, specifically cyclin dependent kinase 5 (CDK5) and GSK-3β, has been evaluated in previous studies as important molecular targets in treating AD. Non-selective CDK5 inhibitors like (R)-roscovitine and (R)-CR8 are still under investigation to provide better understanding of their neuroprotective effect [134,135]. Likewise, different classes of inhibitors of GSK3β such as lithium ions, thiazoles, indirubins, thiadiazolidinones, hymenialdisine and others have been reported for their potential protective effect against AD [136,137,138,139,140]. Additionally, immunological approaches (active and passive immunization) against various forms of tau are areas of increasing research interest. Unlike anti-amyloid antibodies and vaccines, anti-tau vaccine and antibodies are reported to have promising effects against AD [141,142]. These strategies are expected to reduce the formation of tau tangles and help in synapsis and neuronal plasticity.

4. Prospect of APOE4 as a Drug Target for AD

Another independent risk factor in AD is apolipoprotein E4 (APOE4) protein, which normally helps in the transportation of the cholesterol through the APOE receptors [143]. Higher expression of APOE4 has been reported to be associated with the late onset of the disease [144]. There is evidence that APOE4 proteins induce Aβ aggregation and reduce Aβ clearance [145]. Furthermore, APOE4 proteins not only target Aβ interaction, but are also linked to tau hyperphosphorylation, energy metabolism and inflammation in neurons [146,147,148]. The inflammatory response in the brain leads to the proteolysis of APOE4 that may lead to the formation of highly bioactive toxic molecules [149]. Formation of these bioactive fragments of the APOE4 disturb the energy metabolism by altering mitochondria. Furthermore, early evidence shows that APOE4 effects are more pronounced in females, implying possible participation of sex hormones such as estrogen in determination of AD progression. Studies to unravel the actual cause of the gender effect could be a guide for novel approaches to prevent AD [150].

5. Reactive Oxygen and Reactive Nitrogen Species in AD

Oxidative stress accumulation is an important event during AD that worsens as the disease progresses [151]. Oxidative stress is triggered by the accumulation of free radicals such as reactive oxygen species (ROS) and reactive nitrogen species (RNS) and the inability of the cells to clear these reactive molecules. Formation of these free radicals occurs in the electron transport chain due to the loss of the electrons during transfer in the mitochondrial membrane [152]. Formation of free radicals can occur in neurons by multiple factors including mitochondrial dysfunction, impaired autophagy, disruption of lipid homeostasis, formation of lipofuscin, Aβ-induced oxidative damage and accumulation of transition elements (such as iron, copper, zinc, aluminium and mercury) [152,153,154]. These free radicals can oxidize proteins, lipids and DNA affecting various important metabolic processes in the neuronal cells [153,155]. They can also activate the expression of pro-inflammatory markers, such as NFκB and cytokines, which contributes to the recognition of damaged cells [156]. Clearance of these toxic species (ROS and RNS) and balancing the redox state is a requirement for cells to function normally. Several antioxidant genes are expressed to protect the cells from oxidative damage. Young cells function efficiently to clear these free radicals, whereas older cells are thought to be less efficient in doing so [157]. Accumulation of ROS and RNS for a longer duration puts the cells under chronic oxidative stress and initiates abnormal changes. It is still not clear whether oxidative stress accumulation that occurs during ageing is the cause of AD or the aftermath of the disease progression. No matter what comes first, it is evident that oxidative damage is the detrimental event in AD that kills the neuronal cells [158]. Studies on Aβ42 expressed in yeast show that Aβ42 can cause the mitochondrial dysfunction, enhance stress response and upregulate expression of protective antioxidant genes signifying oxidative stress accumulation [154,159]. Biometals involved in oxidative stress management in the cells include iron, aluminium, mercury, zinc and copper [160]. Altered levels of iron, zinc, copper and aluminium have been reported in AD brains [3,161]. With excessive oxidative stress accumulation, protein degradation by cathepsins in lysosomes may also get impaired due to the formation of oxidizing complex molecules like lipofuscin. This can lead to the impairment of the autophagic clearance. In the meantime, lipofuscin further increases oxidative damage to cells by catalysing the Fenton reaction accelerating formation of free radicals [162].
In early stages of AD, Aβ’s entry in mitochondria disrupts the mitochondrial function and generates free radicals [163]. Additionally, APP and Aβ are also reported to be localized in the membrane of mitochondria thereby disrupting the normal electron transport chain. The disruption causes the loss of electrons from the mitochondrial membrane [164]. In summary, redox dyshomeostasis in cells negatively impacts the cellular processes and metabolism that includes impairments to: protein clearance, mitochondrial function, biometal homeostasis, calcium homeostasis, inflammatory responses and antioxidant capacity [165]. Considering these facts, the search for drugs targeting the early relief from oxidative stress in the neuronal cells could be beneficial for preventing neurodegenerative diseases including AD. Antioxidants extracted from various plants can be a natural source of nutraceuticals and prospective therapeutics. Approaches of antioxidant therapy using natural compounds, such as stilbenes, flavonoids, epicatechin, Gingko biloba extracts, ascorbic acid, melatonin and curcumin, have been found to have beneficial effects against AD [166,167,168,169,170,171,172]. Studies have also shown that the reduction of Aβ in neuronal cells can be achieved using antioxidant therapy. In addition, therapies that restore the normal mitochondrial function or mitochondrial regeneration are also found to restore redox balance in cells [158]. Similarly, ongoing studies of metal chelators in combination with other strategies are also considered as a more effective approach [173,174,175].

6. Single Target Strategies in Management of AD

New approaches to treat AD have also considered other targets that are associated with progression of AD. Regulating γ-amino butyric acid (GABA) receptors is one such approach, as GABA is produced by decarboxylation of the neurotransmitter glutamate which ultimately affects the excitotoxic pathway [176]. There are two different isoforms of GABA receptors: GABA 1 and GABA 2 [177]. Many compounds targeting both receptors have been studied to assess their effectiveness to treat AD, but none have shown promising results [178,179]. Despite the initial unsuccessful clinical trials, there is room for hope and studies are continuing.
Phosphodiesterase is another drug target used in previous studies. It normally cleaves phosphodiester bonds in the secondary signaling molecules such as cGMP and cAMP, thus affecting the signal transduction [180]. Various phosphodiesterase inhibitors have been studied to assess their neuroprotective effect [181,182,183]. Cyclooxygenases, COX-2 in particular, have also been used as a target for treatment of AD. COX-2 has been reported to induce Aβ42 formation by increasing γ-secretase activity through prostaglandin formation [184]. Furthermore, COX-2 has been found to activate NMDA receptors, thereby causing excitotoxicity in neuronal cells [185]. Recent studies have shown limited involvement of COX-2 in Aβ deposition, however COX inhibitors are still beneficial in treatment of AD [186].
Histaminic receptors (H3) have a role in releasing neurotransmitters such as acetylcholine, dopamine, nor-epinephrine, histamine and serotonin [187]. Interference in functions of these receptors by antagonists has revealed that they have a protective role in tau-associated memory deficits [188]. Serotonergic receptors have also been studied for their role in cognitive dysfunction, amyloid formation and neuroinflammation during progression of AD [189]. Inhibition of these receptors activated neurotransmitters, glutamate and acetylcholine in particular, improving cognition in AD patients and suggesting that these receptors can be important targets for drug development [190].
The peroxisome proliferator-activated receptor γ (PPARγ) is another target in drug development against AD. PPARγ, a nuclear receptor found in restricted brain areas, has important role in glucose and lipid metabolism [191,192]. Furthermore, PPARγ has been demonstrated to enhance neuronal inflammation and damage [193]. Inhibition of these receptors reduced Aβ aggregates and expression of neuroinflammatory mediators [194,195]. Agonists of PPARγ were found to have other functions in AD brains including clearance of Aβ, disaggregation of Aβ plaques, and reduction of APOE4 expression, however severe side effects of these compounds led to the cessation of the clinical trials [196,197,198,199,200].
The endocannabinoid system is another pathway that is targeted for drug development for AD. Targeting different processes in the system has been reported to increase cognition and the anti-inflammatory response. Targeting the endocannabinoid system reduced Aβ-induced toxicity and tau hyperphosphorylation [201]. The system comprises two lipid molecules derived from endogenous arachidonic acid which bind with two different receptors, CB1 and CB2. The binding of these lipids depends on two different hydrolytic enzymes, fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL) [202]. Drug designers have considered these molecules, including the receptors and enzymes involved, as potential therapeutic target [203,204,205].
Cholesterol has been considered as one of the risk factors for AD and it has been demonstrated that it also contributes to the formation and accumulation of Aβ [206]. Cholesterol lowering drugs are thus important drugs that show benefits against AD [207]. While some studies of statins and AD are inconclusive, other studies support their neuroprotective role [208,209]. Studies in yeast show that statins reduced the levels of intracellular recombinant Aβ implying the possible induction of autophagy [210].
Neurotrophic factors or neurotrophins (NTs) including nerve growth factor (NGF), brain-derived neurotrophic factor (BDNF), neurotrophin 3 (NT3), and neurotrophin 4 (NT4) are crucial for development, maintenance, repair and survival of neuronal populations [211,212,213]. These polypeptides exert their actions through binding and specifically activating tropomyosin receptor kinases (Trk) of either TrkA, TrkB and TrkC [214]. The activation of the receptors induces phosphorylation of the cytoplasmic domain kinases and stimulates signaling pathways including phosphatidyloinositol-3-kinase (PI3K)/protein kinase B (Akt), mitogen-activated protein kinase (MAPK) and phospholipase C-γ1 [215] which are responsible for survival, growth, neuronal differentiation, neurogenesis and neuroplasticity [214,216].
First discovered by Levi-Montalcini in 1951 [217], NGF was shown to be important in the neuronal plasticity and survival of cholinergic neurons in the cerebral cortex, hippocampus, basal forebrain and hypothalamus [218]. The reduction in NGF amount and activity are substantial in the AD [219,220,221]. Therefore, administration of NGF to targets survival and synaptic functions of cholinergic neurons could be useful in the therapeutic prevention and treatment of the disease [222,223].
Exogenous administration of NGF in animals was found to improve the cholinergic system in the CNS, particularly in the forebrain and hippocampus, leading to enhanced cognitive function [224,225,226,227,228]. However, exogenous administration of NGF to combat AD is a difficulty as this protein does not normally pass through the blood brain barrier (BBB) [229,230,231]. In addition, direct delivery of neurotrophic factors may exert serious peripheral side effects [232]. These limitations have brought about innovations to enhance the bio-delivery of NGF for AD therapy by using stem cells, viral vectors, small molecule modulators and most recently, encapsulated cell biodelivery (ECB) [233]. While these methods are known to be costly, hard to administer and precarious, consumption of bioactive compounds from natural products are increasingly preferred in an attempt to slow down and prevent disease progression.

7. Drug Combinations as a Strategy for AD Therapy

Requirements of multifactorial drug design resulted in the development of drug combination strategies. Drugs that target at least two molecular targets of AD were tested for efficacy against AD. A series of hybrid compounds produced by combining two drugs that inhibit AChE and BACE1 has been reported [234,235,236,237,238,239]. Other combinations were found to be effective in metal chelating activity and antioxidant properties with less toxicity [237,239]. Similarly, combinations of AChE with GSK3β inhibitors, MAO inhibitors, metal chelators, NMDAR inhibitors, 5-HT receptor inhibitors, histaminic receptors inhibitors and phosphodiesterase inhibitors have been studied [240]. Some drugs designed in this way have been reported to alleviate AD. However, a number of drug combinations were discontinued due to their adverse effects or low activity [240]. Furthermore, combinations of BACE1 with a GSK3β inhibitor, metal chelators with MAO-B and phosphodiesterase inhibitors were also studied for their efficacy as multitarget therapy [241,242,243,244]. The different combinations of drugs, targeting multiple events of AD pathology hallmarks, may provide substantial protection and possibly cure AD (reviewed in [240]).

8. Restoring Protein Homeostasis as a Novel Multifactorial Approach

Disruption of protein homeostasis is one of the major hallmarks of the age-related neurological disorders [245]. There are different mechanisms by which proteostasis is regulated within the cell. The unfolded protein response (UPR), ubiquitin proteasome system and autophagy are responsible for maintaining the protein balance within cells [246]. Impairment in these processes leads to the accumulation of unwanted cytosolic garbage. Ageing normally comes with less efficient cellular processes including proteostasis [247].
The UPR occurs as misfolded proteins start accumulating, causing ER stress. The inositol response element 1 (IRE1), activating transcription factor 6 (ATF6) and PRK-like ER kinase (PERK) proteins play crucial roles in sensing the presence of aberrant proteins and triggering the upregulated expression of chaperones and foldases to rectify protein folding errors. This ultimately takes the aberrant proteins through ER-associated proteasomal degradation. During ageing, the proteins involved in the UPR are expressed in low levels signifying that upregulating the expression of the proteins will be a potential strategy for preventing or slowing down protein misfolding diseases during ageing [248].
During post translational modification of proteins, ubiquitination of lysine residues is normal. This allows the selective degradation of inappropriately folded proteins mitigating their negative effects. Three different enzymes, namely E1 activating/carrier ubiquitin enzyme, E2 and E3 ligase, interact to transfer ubiquitin to the target proteins’ lysine residues. The ubiquitin tags are removed by deubiquitinating enzymes in normal conditions. But in the case of misfolded proteins, the process of ubiquitination continues several times leading to the formation of polyubiquitin tags in the protein, which is recognized by proteasomal receptors for further processing. Proteasomes are found as complexes called 26S complex that contain two subunits (20S catalytic unit and 19S regulatory unit). The catalytic 20S unit is composed of three proteolytic subunit classes β1 (caspase like activity), β2 (trypsin like activity) and β5 (chymotrypsin like activity) [249]. These proteases not only target polyubiquitinated proteins but also degrade the oxidized proteins [246]. This process of tagging the unwanted proteins with polyubiquitin tags and degrading them through the proteasomes is also referred to as the ubiquitin proteasome system.
Clearance of misfolded proteins, damaged organelles and global turnover of the components of the cell takes place through autophagy [250]. Autophagy can be of three different types including microautophagy, chaperone mediated autophagy and macroautophagy [250]. Microautophagy is the normal process of engulfment of unwanted material of the cytosol in the lysosomal vesicle [251]. In the lysosomal vesicle, different enzymatic action degrades the engulfed unwanted material. Chaperone-mediated autophagy is another system which acts through chaperone proteins (heat shock proteins like Hsp70), which initially bind with the misfolded protein and refold it. When the refolding fails, the chaperones drive these bound materials to the lysosomal vesicles through the lysosomal receptor (LAMP2A) for lysosomal degradation [252]. In addition to these local events of protein clearance, a huge turnover of the cellular molecules/organelles occurs through macroautophagy for supply of required components during different stages in the cell cycle [253]. Macroautophagy, also termed as autophagy hereafter, was initially identified as the effect of starvation [254]. Increase in AMP/ATP ratio during starvation activates AMPK and inhibits protein kinase B (Akt)/mechanistic target of rapamycin (mTOR) pathway, activating the initiation of the autophagosome formation. ROS (dihydronicotinamide-adenine dinucleotide phosphate/NADPH oxidase-induced) accumulation, PI3K/Akt/mTOR inhibition, AMPK, Beclin1, transcription factor EB (TFEB) and sirtuin 1 (SIRT1)/fork head box like protein (FOXO) activation are known pathways for inducing autophagy [250,255,256,257,258].
In AD, accumulation of aberrant Aβ is an example of the disruption of protein homeostasis. Disruption of proteostatis is considered to be the major cause of Aβ accumulation. Mitochondrial dysfunction, ROS accumulation, lipid peroxidation and expression of stress response genes are the consequence of Aβ toxicity in cells. Alterations in the redox state, impairment in protein degradation system, altered distribution of biometals, cellular senescence and cell death are the consequences of the impact in neuronal cells [2,3]. Furthermore, generation of lipofuscin due to increased oxidative stress is another part of the story as these highly lipophilic reactive species catalyse the Fenton reaction causing generation of more free hydroxyl radicals. This leads to the irreversible damage of the cells by oxidizing lipids, proteins and DNA [153]. Impairment in lysosomal and proteasomal degradation is also associated with accumulated lipofuscin in these cellular compartments [259]. Lipofuscin is a complex of molecules formed by the combination of lipid peroxides, oxidized proteins, transition metals and some carbohydrates [153]. Disrupted autophagy may also result in the impairment in lipolysis causing the lipid dyshomeostasis in the cells [260,261]. In intracellular environments of dividing cells, lipofuscin is neither digested nor exocytosed, however it is diluted through cell division. Conversely in neuronal cells, lipofuscin aggregates cannot be diluted through cell division as neuronal cells remain in the G0 part of the cell cycle. Attempted division of these cells induces cell death [153,262]. The drug that clears lipofuscin from the cell could restore protein homeostasis and possibly cure AD. Overall, protein homeostasis maintenance and redox state balance in cells could provide efficient early intervention and limit the disease progression. Targeting the restoration of protein homeostasis has also been hypothesized to provide protection against various other neurodegenerative diseases.
Restoring the protein balance is believed to protect neuronal cells to overcome age-related changes. Protein dyshomeostasis is considered to be a prime factor of oxidative damage, mitochondrial dysfunction, epigenetic alterations, altered biometal distribution, accumulation of aberrant proteins, aggregation of proteins, lipid dyshomeostasis, altered energy metabolism and cell death during progression of AD. Furthermore, inducing processes like autophagy may even increases synapsis, cognition and longevity of the neuronal cells [263,264]. These multiple effects of restoring protein balance in ageing cells will reduce the burden of the neurodegenerative disease.

9. Multiple Targets of Polyphenols against AD

Polyphenols are a class of compounds that are commonly found in many plants. Four major classes of polyphenols including flavonoids, stilbenes, phenolics and lignans are highly regarded as potential therapeutics for neurodegeneration, cardiovascular diseases, cancer and obesity. Many more polyphenolic compounds are yet to be studied for their potency in AD and other neurodegenerative diseases. Polyphenols are classified according to their structure (reviewed in [265]). Structures of some important polyphenols that are described in the text are depicted in Figure 2.
Polyphenolic compounds abound in mushrooms and are one of their main antioxidants. They are mainly phenolic acids which can be divided into groups of either hydroxybenzoic acids and hydroxycinnamic acids derived from the non-phenolic molecules benzoic and cinnamic acid, respectively [266]. The most common benzoic acid derivatives present in mushrooms were reported as p-hydroxybenzoic, protocatechuic, gallic, gentisic, homogentisic, vanillic, 5-sulfosalicylic, syringic, ellagic and veratric acids as well as vanillin. Meanwhile, cinnamic acid derivatives mainly found in mushrooms were p-coumaric, o-coumaric, caffeic, ferulic, sinapic, 3-o-caffeoylquinic, 4-o-caffeoylquinic, 5-o-caffeoylquinic and tannic acids [266].
It is known that only plants synthesize flavonoids, while animals and fungi are not capable of it. However, accumulating studies indicate the presence of flavonoids in different edible mushrooms [267]. The presence of flavonoids in mushrooms could arise from absorption from the substrates where they grow or from neighboring plants by establishing symbiotic interactions via formation of mycorrhizae [268].

9.1. Polyphenols as Antioxidants

Naturally occurring polyphenols provide protection against neurodegeneration through their role as antioxidants [269]. Dietary polyphenols have direct ROS scavenging activity [270]. Several polyphenolic antioxidants identified in common edible mushrooms include protocatechuic acid, p-coumaric, and ellagic acid as well as gallic acid, pyrogallol, homogentisic acid, 5-sulfosalicylic acid, chlorogenic acid, caffeic acid, ferulic acid and quercetin [271,272]. Most of these polyphenols donate electrons to the free radicals thus neutralizing them, which ultimately reduces the levels of ROS within the cells. Polyphenols activate Nuclear factor erythroid 2-related factor 2 (Nrf2), a basic leucine zipper transcription factor. Nrf2 normally is complexed with Kelch-like ECH-associated protein 1 (Keap1) in the cellular environment inhibiting Nrf2′s nuclear translocation. Furthermore, Keap1 also facilitates ubiquitination and proteasomal degradation of Nrf2 [273]. The separation of Nrf2 from Keap1 leads to activation and nuclear translocation of Nrf2, where it complexes with musculoaponeurotic fibrosarcoma (Maf) proteins. This heteromeric Nrf2-Maf complex then binds with antioxidant response element (ARE) sequences located upstream to the phase II detoxifying genes upregulating their expression. Phase II antioxidant genes encode proteins, such as heme oxygenase 1, γ-glutamyl cysteine synthetase, peroxiredoxins, glutathione reductases, thioredoxin reductase, drug metabolizing and detoxification enzymes NAD(P)H quinone dehydrogenase 1, glutathione-S-transferase, uridine diphosphate-glucuronosyltransferase and regulators, transketolase, PPARγ-coactivator 1 β (PGC1-β), etc [274]. These proteins act in the cell as antioxidant proteins, having a major role in restoration of the redox imbalance and cellular signaling [275,276]. Additionally, polyphenols also elucidate their antioxidant property through inhibition of NADPH oxidase (NOX) activities [277]. NOX proteins are transmembrane proteins that signal the immune modulators through ROS generation [278]. Lower levels of ROS may be important for cellular signaling, however, at higher levels they can cause damage to the neuronal cells. These proteins, found to be involved in increasing Aβ-induced oxidative stress, could be potential therapeutic targets for AD [279].
Oxidative damage is more prominent when the damage is coupled with mitochondrial dysfunction. Enzymes such as monoamine oxidases (e.g., MaoB) increase the cellular stress by producing hydrogen peroxide [280]. In brains, monoamine oxidase activity of substrate neurotransmitters causes mitochondrial damage, while dietary polyphenols have been found to inhibit MaoB, thus decreasing the ROS generation and mitochondrial dysfunction [36]. Additionally, polyphenols also aid in regeneration of mitochondria in the cells through activation of the master regulator SIRT1 [281]. SIRT1 is a NAD+-dependent histone deacetylase enzyme that has multiple targets for deacetylation. SIRT1′s involvement in reducing oxidative stress comes from deacetylation of its substrate PGC-1α, which activates nuclear respiratory factors (Nrf1 and Nrf2) and peroxisome proliferator-activated receptor (PPARα) [282]. Further downstream, these molecules enhance the expression of transcription factor A, mitochondrial (TFAM) that initiates the transcription and replication of mitochondrial DNA ultimately causing the regeneration of mitochondria [282]. The activation AMPK, either directly or indirectly (through SIRT1 activation) activates PGC-1α, thus helping in mitochondrial biogenesis.
Biometals such as iron and copper are the major contributors of ROS formation in defunct mitochondria [283]. Quercetin, baicalein, curcumin, etc., are found to provide a protective antioxidant property also through biometal chelation [4,284,285]. Furthermore, alterations in biometal distribution in the neuronal cells is also an important hallmark of AD. The mechanism through which polyphenols act as antioxidants in the cellular environment is schematically presented in Figure 3. Antioxidants can also act as pro-oxidants in certain sub-optimal concentrations and cause oxidative damage to the cells. Thus, their optimum concentration needs to be considered prior to their application.

9.2. Modulation of Protein Homeostasis and Longevity with Polyphenols

Dietary polyphenols modulate the protein quality control mechanisms increasing the cellular efficiency to clear misfolded proteins. Apart from induction of autophagic clearance, the UPR and ubiquitin proteasome system are also modulated by dietary polyphenols [286,287,288]. The ability of polyphenols to activate lysosomal biogenesis and increase longevity make them an important class of neuroprotective compounds [5,34,37]. In addition, some of the polyphenols like EGCG and curcuminoids reduced the lipofuscin granules in cells, which normally are impossible to degrade or exocytose from the cell [15,289]. Reduction of lipofuscin in the cell can contribute to the restoration of the protein homeostasis by reducing the damage to autophagosomes and proteasomes.
Most of the polyphenolic compounds act through upregulation of the expression of the master regulator SIRT1 [290]. The SIRT1 protein has been found to have multiple targets that play a vital role in regulating major cellular processes (refer to Figure 4) [290]. The activation of AMPK/Unc-51 like autophagy activating kinase 1 (ULK1), transcription factor EB (TFEB), Fork head box O transcription factors (FOXO), deacetylation of p53 and inhibition of PI3K/Akt/mTOR, NFkB, MAPK and the c-Jun N-terminal kinases (c-JNK) pathway are important cellular processes that will induce autophagy through SIRT1 [291,292,293]. Most of these molecular targets are deacetylation substrates of SIRT1. Activation of transcription factors like TFEB reinforces the cellular autophagy by activating lysosomal biogenesis. TFEB itself is another master regulator for the coordinated lysosomal expression and regulation (CLEAR) network. The CLEAR network has important roles in various cellular processes. Energy metabolism, DNA metabolism, steroid biosynthesis, protein clearance, haemoglobin degradation, antigen presentation, phagocytosis and signal transduction are important events regulated by TFEB [294,295].
Similarly, SIRT1 has a significant role in determining cellular fate via Fork head transcription factors (FOXO1 and FOXO3). The deacetylated form of these transcription factors are major contributors of autophagy activation, cell cycle arrest, stress resistance (expression of manganese superoxide dismutase) and immune modulation. Reduction in the levels of FOXO by ubiquitination and proteasomal degradation with the help of SIRT1 reduces the levels of acetylated forms. Reduction in acetylated FOXO’s suppresses cell death caused by apoptosis driving cells towards survival and increasing longevity (refer to Figure 5) [292,296]. This is of particular interest for neurodegenerative diseases, where survival of neuronal cell after damage is crucial. It has been illustrated that polyphenols activate these master regulators of longevity (Nrf2, SIRT1 and AMPK) providing unprecedented protection against various disease [276,297,298]. However, limited bioavailability of these dietary polyphenols in human has limited their application. Polyphenols such as hydroxytyrosol, oleuropein aglycone, curcumin, resveratrol, rotenone, rutin, myricetin, urolithin A, epigallocatechin 3-gallate (EGCG), ferulic acid, genipin, etc. have been reported to induce autophagy. The olive oil polyphenol, hydroxytyrosol activates AMPK pathway and is reported to reduce Aβ levels in mouse models of AD [28,299]. Similarly, oleuropein aglycone has been reported to activate SIRT1/AMPK/mTOR and TFEB mediated autophagy [300,301].
Curcumin, one of the most studied polyphenols, has multifactorial benefits in balancing the protein homeostasis by activation of AMPK/ULK1 and inhibition of PI3K/Akt/mTOR through activation of SIRT1 [38]. EGCG, a catechin family polyphenol, inhibits the suppressors (Bcl2 and Bcl-XL) of Beclin1. However, the activity of this polyphenol is also dependent on the concentration of the compound. A higher concentration of EGCG inhibits autophagy and induces apoptosis, whereas, lower concentrations induce autophagy that also degrade lipid droplets through a Ca2+/CAMKKB/AMPK dependent mechanism. Thus, the concentration of polyphenols is a crucial factor before considering it as a therapeutic option. EGCG has also been reported to reduce the catalytic activity of 19S and 20S proteasomal proteins, deactivate NFkB pathway and enhance p53 tumour suppressor protein expression [302]. An important feature of EGCG also includes its ability to inhibit lipofuscin formation, which otherwise impairs autophagy and the proteasome during ageing [15].
Resveratrol is another important polyphenol frequently studied for its beneficial effect in increasing longevity and balances cellular protein homeostasis. The activation of SIRT1/AMPK and extracellular signal-regulated kinases (ERK1/2) is the molecular mechanism by which this polyphenol was found to be neuroprotective [9,303,304]. The metabolite of ellagitannin, urolithin A, extracted from pomegranate has been reported to activate autophagy through SIRT1 activation [305]. Furthermore, the natural compound was also found to increase mitophagy and longevity in a Caenorhabditis elegans (C. elegans) model that has provided insight on human neurodegeneration [49]. Quercetin has shown multiple benefits in human health by enhancing autophagy through SIRT1activation, inhibiting proteasomal degradation (inhibition of all the catalytic subunits), reducing proliferation and activating apoptosis [306]. Apart from autophagy inducers, hesperitin and hesperidin have also been reported to have negative effects on Aβ-induced autophagy and glucose metabolism impairment [307,308].

9.3. Polyphenols and Cellular Lipid Balance

Polyphenols are also considered as potential therapeutic agents against obesity and other life-threatening conditions [309,310,311]. This property of polyphenols is associated with the activation of AMPK, which targets lipid metabolism as well [312]. Activation of AMPK decreases the activity of acetyl CoA carboxylase, HMG-CoA reductase and diacylglycerol acyl transferase, and thus avoids hepatic accumulation of lipids [313,314]. These actions of AMPK reduce the levels of free fatty acids as well as the complex lipids. Polyphenols are also found to inhibit the adipogenesis by inhibiting proteins like PPARγ [315,316]. Additionally, as explained in previous sections, polyphenols increase autophagic clearance. Induction of autophagy is not only limited to restoring the protein balance but is also associated with the degradation of lipids to meet the energy demands of the cells. Thus, polyphenols can also reduce lipid accumulation in the intracellular environment [260]. AD is also termed as Type III diabetes due to its similarity with diabetes. High levels of cholesterol have been found to be associated with AD brains [317]. Lowering the levels of cholesterol has been an important approach for the treatment of AD, despite limited success. Furthermore, studies support increased activity of γ-secretase and β-secretase with higher levels of lipids in the membrane environment that could contribute to increased Aβ levels in the brain [210]. Considering these facts, polyphenols are hypothesized to have their neuroprotective action in part through the restoration of lipid homeostasis.

9.4. Anti-inflammatory Activity of Polyphenols

ROS act as signaling molecules for induction and release of pro-inflammatory mediators including NFκB and cytokines. NFκB exists in an inactivated form bound to an inhibitor referred to as p65/p50 dimer in normal conditions [318]. When this complex gets activated by increased ROS, the p65/p50 dimer translocates to the nucleus upregulating expression of the inflammatory markers [319]. The expression of these inflammatory mediators inside the cells triggers the downstream process of inflammation. Deacetylation of NFκB through the action of SIRT1 at specific amino acid residues renders it inactivated and reduces the inflammatory response by reducing the expression of downstream genes [318]. Since polyphenols are antioxidants capable of lowering the ROS in the cells, they can downregulate the expression of proinflammatory mediators [320]. However, the highest anti-inflammatory activity of polyphenols is attributed to their ability to activate the master regulator SIRT1 [321]. Many polyphenols have been reported to have an anti-inflammatory effect which could provide the basis for protection against diseases with chronic neuroinflammation/inflammation.

9.5. Polyphenols as Anti-amyloid Agents

Oleuropein, an olive polyphenol, is found to increase α-secretase activity. Thus, it prevents cells from producing Aβ: instead such activity results in the formation of the Aα peptide [322]. Formation of Aα instead of Aβ is anti-amyloidogenic, which may be helpful in reducing the Aβ-associated toxicity. Some polyphenols (such as rutin) reduce the β-secretase activity [6]. Similarly, other polyphenols disaggregate the amyloid aggregates in vitro [6,323]. Furthermore, the ability of polyphenols to lower the cholesterol levels in cells also favors the reduced activity of β-secretase and γ-secretase [6,317]. Apart from the anti-amyloid functions, polyphenols also possess the ability to inhibit tau aggregation [324].
Through characterization of the cell-free extracts of different bacteria, fungi and yeast, Lee et al. (2007) identified the BACE1 inhibitory effects of different mushrooms [325]. Mushroom species having anti-BACE1 effects were Flammulina velutipes, Pleurotus ostreatus, Grifola frondosa, Dictyophora echinovolvata, Fomitella fraxinea and Inonotus obliquus. Hispidin, a polyphenolic compound found in abundance in the mushroom Phellinus linteus inhibits BACE1 non-competitively and scavenges free radicals [326]. BACE1’s inhibitory effect of Auricularia polytricha has also been indicated to be hispidine mediated [327].

9.6. Polyphenols in Cognition and Synapsis

Polyphenolic compounds like α-isocubebenol, tacrine and their derivatives, 2′,4′-dihydroxy-6′methoxy-3′,5′-dimethyl-dihydrochalcone, tetrahydropyranodiquinolin-8-amines, quercetin and tiliroside have been shown to have neuroprotective properties attributed to their inhibiting activity against acetylcholine esterase [328,329,330,331]. In addition, some other polyphenols, including genistein, luteolin-7-O-rutinoside and silibinin, are reported to have a moderate effect on the butyrylcholine esterase [330]. Among the polyphenols, flavonoids are an important class of polyphenols that have anti-choline esterase activity [167]. Flavonoids extracted from Ginkgo biloba have been reported to have inhibitory effects against acetyl choline esterase [168]. Molecular docking experiments revealed the mechanism of action of quercetin was through strong hydrogen bond formation with certain amino acids of AChE, thus leading to competitive inhibition of AChE. Similarly, macluraxanthone exhibited non-competitive type interference with the activity of acetyl choline esterase [167]. The combination of numerous hydrogen bonds with several amino acids and hydrophobic interaction may be responsible for how these polyphenols inhibit acetylcholine esterase activity [332].
Polyphenols exert neuroprotective effects in experimental systems but there is a need to translate this in guidelines for neuroprotection of aging populations. For translation of animal studies to human trials, dose accuracy plays a critical role. For example, consider resveratrol levels in Table 1: an effective dose in mice is 60 mg/kg/d by oral administration. In humans this translates to ~290 mg for a 60 kg person per day [333]. Such levels are rarely reached. In the case of resveratrol, the suggested daily intake is 200 mg/day and this is unlikely to be a protective level. In addition, alterations in polyphenol administration routes may reduce the amount of polyphenol to be used on daily basis, signifying the benefits of alternative administration strategy. However, long term uptake of the polyphenol could still have beneficial effects in lower doses. On the other hand, some nutraceutical products may contain the polyphenol at more than the optimal amount, which could have negative effects in brain health [334]. This bimodal activity of polyphenols should be highly considered before translating the beneficial effects of the polyphenols for human use.

10. Future Directions

In order to reap the full benefits of polyphenols as therapies in AD, some limitations should be considered—especially in regard to safety, pharmacokinetics, bioavailability, delivery system, administration route, dose efficiency and clinical status (reviewed in [334]). In terms of safety, polyphenols were generally regarded as safe and well-tolerated in animals as well as humans with no notable side effects even for high and repeated dosages [335]. If any, side effects are usually mild, tolerated, and transient: for instance, minor headaches, dizziness, gastrointestinal problems, and skin rashes. Another important point to contemplate is the possible interaction of clinically-prescribed drugs with polyphenols, as polyphenols are currently viewed as nothing more than a supplement, and far from being a substitute for prescription drugs. For example, flavonoids in grapefruit juice demonstrated potent inhibition of the cytochrome P450 (CYP) protein family, critical for drug metabolism. The abrupt inhibition of CYP may potentially lead to excessive buildup of drugs increasing toxicity [336,337,338]. Regardless, polyphenols taken exclusively were harmless either in short, medium, or long-term supplementation in humans [339,340,341,342,343,344,345], which certainly encourages their application. Despite countless attempts proving the AD-ameliorating efficacy of polyphenols in a wide range of in vitro, in vivo, and epidemiological studies, the translation into human trials is indeed difficult, and failure was common in the early stages of most clinical trials [346]. However, their efficacy has improved over time with further modification of multiple factors, including effective dosage and period of administration. As a result, resveratrol [347,348] and Ginkgo biloba (flavone glycosides and terpene lactones) [343,345,349] showed promising results in the initial phases, while EGCG stands out by reaching phase III of clinical trials [350]. There is no doubt about the benefits and potential of polyphenols in the management of AD, but a poor understanding of pharmacokinetics and pharmacodynamics has restricted their applicability. In many instances, their bioavailability in the CNS was limited due to low absorption in the gastrointestinal tract, rapid metabolism, systemic elimination, and impermeability across the BBB [351,352,353]. Processing and first-pass metabolism of these dietary polyphenols, which occurs at different levels, including the stomach, small intestine, large intestine, circulatory system and liver, may cause significant changes in polyphenol structure, quantity and biological activity [354,355]. Furthermore, the gut microbiome also takes part in metabolizing these bioactive compounds [355,356]. Studies suggest only 5-10% of the dietary polyphenols are absorbed, leaving much room for improvements to increase the bioavailability of these potential therapeutics. Even more critical in neurodegenerative disorders is the requirement for these polyphenols to cross the BBB from the bloodstream to the brain tissue to reach their target, which depends on their lipophilicity [357,358]. Hence, future research should be focused on optimizing the bioavailability of these compounds in the human body, particularly in brain tissues, to have enhanced effects. Recent studies involving the encapsulation of these bioactive compounds into stable nanoparticles and microparticles could be significant [359]. The possibilities of administering these compounds through a different route into the human body should be considered: for instance, intranasal or intravenous administration to avoid inactivation during the first-pass metabolism and gut microflora intervention. Improvement in targeted delivery through engineering particles in such a way that their bioavailability is increased would be the basis for further research. Considering these facts improvements made to enhance the bioavailability of curcumin [360,361] and resveratrol [362] were successful to some extent, which provides a roadmap for future studies.

Author Contributions

S.D. conceived, designed and wrote the paper. N.K. and C.W.P. co-drafted and revised the paper. V.S. co-supervised and proof read the article. I.M. and B.A. conceived and supervised the paper.

Funding

The study was supported by RMIT Research Stipend Scholarship and RMIT Tuition Fee Offset Scholarship awarded to Sudip Dhakal.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADAlzheimer’s Disease
Aβ42β-amyloid of 42 amino acids
Amyloid α
NGFNerve Growth Factor
FDAFood and Drug Administration
NMDARN-Methyl-D-Aspartic Receptor
APPAmyloid Precursor Protein
BACEβ-Secretase
NFTNeurofibrillary Tangle
NFκBNuclear factor kappa B
MAPKMitogen-Activated Protein Kinase
GSK3βGlycogen Synthase Kinase - 3β
CDKCyclin dependent kinase
APOEApolipoprotein E
ROSReactive Oxygen Species
RNSReactive Nitrogen Species
GABAγ-amino butyric acid
cGMPCyclic guanosine monophosphate
cAMPCyclic adenosine monophosphate
COXCyclooxygenase
PPARγPeroxisome proliferator-activated receptor γ
FAAHFatty acid amide hydrolase
MAGLMono acyl glycerol lipase
BDNFBrain-derived neurotrophic factor
NTNeurotrophin
TrkTropomyosin receptor kinase
PI3KPhosphatidylinositol-3-kinase
AktProtein kinase B
BBBBlood Brain Barrier
ECBEncapsulated Cell Bio-delivery
AChEAcetylcholine esterase
MAOMonoamine oxidase
UPRUnfolded protein response
IREInositol response element
ATFActivating transcription factor
PERKProtein kinase RNA-like endoplasmic reticulum kinase
LAMPLysosome associated molecular pattern
ATPAdenosine triphosphate
AMPKAdenosine monophosphate kinase
mTORMechanistic Target of Rapamycin
NADPHDihydronicotinamide-adenine dinucleotide phosphate
NOXNADPH oxidase
TFEBTranscription factor EB
SIRT1Sirtuin 1
FOXOFork head box like protein O
NrfNuclear factor erythroid-2 related factor
KeapKelch-like ECH-associated protein 1
MafMasculoaponeurotic fibrosarcoma
AREAntioxidant response element
UDPUridine diphosphate
PGC1PPARγ coactivator-1
TFAMTranscription factor A, mitochondrial
EGCGEpigallocatechin-3-gallate
ULKUnc-51 like autophagy activating kinase
c-JNKc-Jun N-terminal kinase
CLEARCoordinated lysosomal expression and regulation
HDACHistone deacetylase
AtgAutophagy related
CAMKKCalcium/Calmodulin-dependent protein kinase kinase
BclBeclin
ERKExtracellular signal-regulated kinases
HMGCoA3-hydroxy-3-methyl-glutaryl-Coenzyme A
DNMTDNA (cytosine-5)-methyltransferase
HOHeme oxygenase
HSPHeat shock protein
TNFTumor necrosis factor
ILInterleukin
SODSuperoxide dismutase
CREBcAMP response element-binding protein
BaxBeclin-2- associated X

References

  1. Macreadie, I.G.; Arvanitis, C.; Bharadwaj, P. Finding chemopreventatives to reduce amyloid beta in yeast. Neural Regen. Res. 2016, 11, 244–245. [Google Scholar] [CrossRef]
  2. Sharma, P.; Srivastava, P.; Seth, A.; Tripathi, P.N.; Banerjee, A.G.; Shrivastava, S.K. Comprehensive review of mechanisms of pathogenesis involved in Alzheimer’s disease and potential therapeutic strategies. Prog. Neurobiol. 2019, 174, 53–89. [Google Scholar] [CrossRef] [PubMed]
  3. Kim, C.A.; Lim, S.; Kim, K.Y. Metal Ion Effects on Aβ and Tau Aggregation. Int. J. Mol. Sci. 2018, 19, 128. [Google Scholar] [CrossRef] [PubMed]
  4. Xiao, L.; Luo, G.; Tang, Y.; Yao, P. Quercetin and iron metabolism: What we know and what we need to know. Food Chem. Toxicol. 2018, 114, 190–203. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, Y.; Chen, Y.; Shaw, A.M.; Goldfine, H.; Tian, J.; Cai, J. Enhancing TFEB-mediated cellular degradation pathways by the mTORC1 inhibitor quercetin. Oxidative Med. Cell. Longev. 2018, 2018, 5073420. [Google Scholar] [CrossRef] [PubMed]
  6. Jimenez-Aliaga, K.; Bermejo-Bescos, P.; Benedi, J.; Martin-Aragon, S. Quercetin and rutin exhibit antiamyloidogenic and fibril-disaggregating effects in vitro and potent antioxidant activity in APPswe cells. Life Sci 2011, 89, 939–945. [Google Scholar] [CrossRef] [PubMed]
  7. Primikyri, A.; Mazzone, G.; Lekka, C.; Tzakos, A.G.; Russo, N.; Gerothanassis, I.P. Understanding zinc(II) chelation with quercetin and luteolin: A combined NMR and theoretical study. J. Phys. Chem. B 2015, 119, 83–95. [Google Scholar] [CrossRef]
  8. Dosenko, V.E.; Nagibin, V.S.; Tumanovskaya, L.V.; Zagorii, V.Y.; Moibenko, A.A. Effect of quercetin on the activity of purified 20S and 26S proteasome and proteasomal activity in isolated cardiomyocytes. Biomed. Chem. 2007, 1, 40–44. [Google Scholar] [CrossRef]
  9. Chen, Y.; Shi, G.W.; Liang, Z.M.; Sheng, S.Y.; Shi, Y.S.; Peng, L.; Wang, Y.P.; Wang, F.; Zhang, X.M. Resveratrol improves cognition and decreases amyloid plaque formation in Tg6799 mice. Mol. Med. Rep. 2019, 49, 3783–3790. [Google Scholar] [CrossRef]
  10. Pallàs, M.; Casadesús, G.; Smith, M.A.; Coto-Montes, A.; Pelegri, C.; Vilaplana, J.; Camins, A. Resveratrol and neurodegenerative diseases: Activation of SIRT1 as the potential pathway towards neuroprotection. Curr. Neurovas. Res. 2009, 6, 70–81. [Google Scholar] [CrossRef]
  11. Suvorova, I.I.; Knyazeva, A.R.; Pospelov, V.A. Resveratrol-induced p53 activation is associated with autophagy in mouse embryonic stem cells. Biochem. Biophys. Res. Commun. 2018, 503, 2180–2185. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, J.; Zhou, Y.; Mueller-Steiner, S.; Chen, L.F.; Kwon, H.; Yi, S.; Mucke, L.; Gan, L. SIRT1 protects against microglia-dependent amyloid-β toxicity through inhibiting NF-κB signaling. J. Biol. Chem. 2005, 280, 40364–40374. [Google Scholar] [CrossRef]
  13. Konings, E.; Timmers, S.; Boekschoten, M.V.; Goossens, G.H.; Jocken, J.W.; Afman, L.A.; Müller, M.; Schrauwen, P.; Mariman, E.C.; Blaak, E.E. The effects of 30 days resveratrol supplementation on adipose tissue morphology and gene expression patterns in obese men. Int. J. Obes. 2014, 38, 470–473. [Google Scholar] [CrossRef]
  14. Chen, Q.; Ganapathy, S.; Singh, K.P.; Shankar, S.; Srivastava, R.K. Resveratrol induces growth arrest and apoptosis through activation of FOXO transcription factors in prostate cancer cells. PLoS ONE 2010, 5. [Google Scholar] [CrossRef] [PubMed]
  15. Cai, S.; Yang, H.; Zeng, K.; Zhang, J.; Zhong, N.; Wang, Y.; Ye, J.; Tu, P.; Liu, Z. EGCG Inhibited Lipofuscin Formation Based on Intercepting Amyloidogenic β-Sheet-Rich Structure Conversion. PLoS ONE 2016, 11, e0152064. [Google Scholar] [CrossRef] [PubMed]
  16. Qin, J.; Xie, L.P.; Zheng, X.Y.; Wang, Y.B.; Bai, Y.; Shen, H.F.; Li, L.C.; Dahiya, R. A component of green tea, (-)-epigallocatechin-3-gallate, promotes apoptosis in T24 human bladder cancer cells via modulation of the PI3K/Akt pathway and Bcl-2 family proteins. Biochem. Biophys. Res. Commun. 2007, 354, 852–857. [Google Scholar] [CrossRef]
  17. Hyung, S.J.; Detoma, A.S.; Brender, J.R.; Lee, S.; Vivekanandan, S.; Kochi, A.; Choi, J.S.; Ramamoorthy, A.; Ruotolo, B.T.; Lim, M.H. Insights into antiamyloidogenic properties of the green tea extract (-)-epigallocatechin-3-gallate toward metal-associated amyloid-β species. Proc. Natl. Acad. Sci. USA 2013, 110, 3743–3748. [Google Scholar] [CrossRef]
  18. Kim, H.S.; Montana, V.; Jang, H.J.; Parpura, V.; Kim, J.A. Epigallocatechin gallate (EGCG) stimulates autophagy in vascular endothelial cells: A potential role for reducing lipid accumulation. J. Biol. Chem. 2013, 288, 22693–22705. [Google Scholar] [CrossRef]
  19. Zhou, J.; Farah, B.L.; Sinha, R.A.; Wu, Y.; Singh, B.K.; Bay, B.H.; Yang, C.S.; Yen, P.M. Epigallocatechin-3-Gallate (EGCG), a green tea polyphenol, stimulates hepatic autophagy and lipid clearance. PLoS ONE 2014, 9, e87161. [Google Scholar] [CrossRef]
  20. Pacheco, S.M.; Soares, M.S.P.; Gutierres, J.M.; Gerzson, M.F.B.; Carvalho, F.B.; Azambuja, J.H.; Schetinger, M.R.C.; Stefanello, F.M.; Spanevello, R.M. Anthocyanins as a potential pharmacological agent to manage memory deficit, oxidative stress and alterations in ion pump activity induced by experimental sporadic dementia of Alzheimer’s type. J. Nutr. Biochem. 2018, 56, 193–204. [Google Scholar] [CrossRef]
  21. Hwang, Y.P.; Choi, J.H.; Yun, H.J.; Han, E.H.; Kim, H.G.; Kim, J.Y.; Park, B.H.; Khanal, T.; Choi, J.M.; Chung, Y.C.; et al. Anthocyanins from purple sweet potato attenuate dimethylnitrosamine-induced liver injury in rats by inducing Nrf2-mediated antioxidant enzymes and reducing COX-2 and iNOS expression. Food Chem. Toxicol. 2011, 49, 93–99. [Google Scholar] [CrossRef] [PubMed]
  22. Longo, L.; Platini, F.; Scardino, A.; Alabiso, O.; Vasapollo, G.; Tessitore, L. Autophagy inhibition enhances anthocyanin-induced apoptosis in hepatocellular carcinoma. Mol. Cancer Ther. 2008, 7, 2476–2485. [Google Scholar] [CrossRef] [PubMed]
  23. Ullah, I.; Park, H.Y.; Kim, M.O. Anthocyanins Protect against Kainic Acid-induced Excitotoxicity and Apoptosis via ROS-activated AMPK Pathway in Hippocampal Neurons. Cns Neurosci. Ther. 2014, 20, 327–338. [Google Scholar] [CrossRef] [PubMed]
  24. Kim, T.W.; Lee, S.Y.; Kim, M.; Cheon, C.; Ko, S.-G. Kaempferol induces autophagic cell death via IRE1-JNK-CHOP pathway and inhibition of G9a in gastric cancer cells. Cell Death Dis. 2018, 9, 875. [Google Scholar] [CrossRef] [PubMed]
  25. Lin, C.-W.; Chen, P.-N.; Chen, M.-K.; Yang, W.-E.; Tang, C.-H.; Yang, S.-F.; Hsieh, Y.-S. Kaempferol Reduces Matrix Metalloproteinase-2 Expression by Down-Regulating ERK1/2 and the Activator Protein-1 Signaling Pathways in Oral Cancer Cells. PLoS ONE 2013, 8, e80883. [Google Scholar] [CrossRef]
  26. Huang, W.W.; Tsai, S.C.; Peng, S.F.; Lin, M.W.; Chiang, J.H.; Chiu, Y.J.; Fushiya, S.; Tseng, M.T.; Yang, J.S. Kaempferol induces autophagy through AMPK and AKT signaling molecules and causes G 2/M arrest via downregulation of CDK1/cyclin B in SK-HEP-1 human hepatic cancer cells. Int. J. Oncol. 2013, 42, 2069–2077. [Google Scholar] [CrossRef]
  27. Qiu, W.; Lin, J.; Zhu, Y.; Zhang, J.; Zeng, L.; Su, M.; Tian, Y. Kaempferol modulates DNA methylation and downregulates DNMT3B in bladder cancer. Cell. Physiol. Biochem. 2017, 41, 1325–1335. [Google Scholar] [CrossRef]
  28. De Pablos, R.M.; Espinosa-Oliva, A.M.; Hornedo-Ortega, R.; Cano, M.; Arguelles, S. Hydroxytyrosol protects from aging process via AMPK and autophagy; a review of its effects on cancer, metabolic syndrome, osteoporosis, immune-mediated and neurodegenerative diseases. Pharmacol. Res. 2019, 143, 58–72. [Google Scholar] [CrossRef]
  29. Zheng, A.; Li, H.; Xu, J.; Cao, K.; Li, H.; Pu, W.; Yang, Z.; Peng, Y.; Long, J.; Liu, J.; et al. Hydroxytyrosol improves mitochondrial function and reduces oxidative stress in the brain of db/db mice: Role of AMP-activated protein kinase activation. Br. J. Nutr. 2015, 113, 1667–1676. [Google Scholar] [CrossRef]
  30. Zrelli, H.; Matsuoka, M.; Kitazaki, S.; Zarrouk, M.; Miyazaki, H. Hydroxytyrosol reduces intracellular reactive oxygen species levels in vascular endothelial cells by upregulating catalase expression through the AMPK-FOXO3a pathway. Eur. J. Pharmacol. 2011, 660, 275–282. [Google Scholar] [CrossRef]
  31. Wang, W.; Jing, T.; Yang, X.; He, Y.; Wang, B.; Xiao, Y.; Shang, C.; Zhang, J.; Lin, R. Hydroxytyrosol regulates the autophagy of vascular adventitial fibroblasts through the SIRT1-mediated signaling pathway. Can. J. Physiol. Pharmacol. 2018, 96, 88–96. [Google Scholar] [CrossRef] [PubMed]
  32. Zhao, B.; Ma, Y.; Xu, Z.; Wang, J.; Wang, F.; Wang, D.; Pan, S.; Wu, Y.; Pan, H.; Xu, D.; et al. Hydroxytyrosol, a natural molecule from olive oil, suppresses the growth of human hepatocellular carcinoma cells via inactivating AKT and nuclear factor-kappa B pathways. Cancer Lett. 2014, 347, 79–87. [Google Scholar] [CrossRef] [PubMed]
  33. Priore, P.; Siculella, L.; Gnoni, G.V. Extra virgin olive oil phenols down-regulate lipid synthesis in primary-cultured rat-hepatocytes. J. Nutr. Biochem. 2014, 25, 683–691. [Google Scholar] [CrossRef] [PubMed]
  34. Miceli, C.; Santin, Y.; Manzella, N.; Coppini, R.; Berti, A.; Stefani, M.; Parini, A.; Mialet-Perez, J.; Nediani, C. Oleuropein aglycone protects against MAO-a-induced autophagy impairment and cardiomyocyte death through activation of TFEB. Oxidative Med. Cell. Longev. 2018, 2018, 8067592. [Google Scholar] [CrossRef] [PubMed]
  35. Rigacci, S.; Miceli, C.; Nediani, C.; Berti, A.; Cascella, R.; Pantano, D.; Nardiello, P.; Luccarini, I.; Casamenti, F.; Stefani, M. Oleuropein aglycone induces autophagy via the AMPK/mTOR signalling pathway: A mechanistic insight. Oncotarget 2015, 6, 35344–35357. [Google Scholar] [CrossRef] [PubMed]
  36. Rajeswari, A.; Sabesan, M. Inhibition of monoamine oxidase-B by the polyphenolic compound, curcumin and its metabolite tetrahydrocurcumin, in a model of Parkinson’s disease induced by MPTP neurodegeneration in mice. Inflammopharmacology 2008, 16, 96–99. [Google Scholar] [CrossRef]
  37. Zhang, J.; Wang, J.; Xu, J.; Lu, Y.; Jiang, J.; Wang, L.; Shen, H.M.; Xia, D. Curcumin targets the TFEB-lysosome pathway for induction of autophagy. Oncotarget 2016, 7, 75659–75671. [Google Scholar] [CrossRef] [Green Version]
  38. Shakeri, A.; Cicero, A.F.G.; Panahi, Y.; Mohajeri, M.; Sahebkar, A. Curcumin: A naturally occurring autophagy modulator. J. Cell. Physiol. 2019, 234, 5643–5654. [Google Scholar] [CrossRef]
  39. Saljoughian, M. Curcumin: A promising antiamyloidogenic agent. U.S. Pharm. 2011, 36, 27–32. [Google Scholar]
  40. Zheng, J.; Cheng, J.; Zheng, S.; Feng, Q.; Xiao, X. Curcumin, a polyphenolic curcuminoid with its protective effects and molecular mechanisms in diabetes and diabetic cardiomyopathy. Front. Pharmacol. 2018, 9, 472. [Google Scholar] [CrossRef]
  41. Yang, F.; Lim, G.P.; Begum, A.N.; Ubeda, O.J.; Simmons, M.R.; Ambegaokar, S.S.; Chen, P.; Kayed, R.; Glabe, C.G.; Frautschy, S.A.; et al. Curcumin inhibits formation of amyloid β oligomers and fibrils, binds plaques, and reduces amyloid in vivo. J. Biol. Chem. 2005, 280, 5892–5901. [Google Scholar] [CrossRef] [PubMed]
  42. Mishra, P.; Paital, B.; Jena, S.; Swain, S.S.; Kumar, S.; Yadav, M.K.; Chainy, G.B.N.; Samanta, L. Possible activation of NRF2 by Vitamin E/Curcumin against altered thyroid hormone induced oxidative stress via NFĸB/AKT/mTOR/KEAP1 signalling in rat heart. Sci. Rep. 2019, 9, 7408. [Google Scholar] [CrossRef] [PubMed]
  43. Sun, Q.; Jia, N.; Wang, W.; Jin, H.; Xu, J.; Hu, H. Activation of SIRT1 by curcumin blocks the neurotoxicity of amyloid-β25–35 in rat cortical neurons. Biochem. Biophys. Res. Commun. 2014, 448, 89–94. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, C.; Zhang, X.; Teng, Z.; Zhang, T.; Li, Y. Downregulation of PI3K/Akt/mTOR signaling pathway in curcumin-induced autophagy in APP/PS1 double transgenic mice. Eur. J. Pharmacol. 2014, 740, 312–320. [Google Scholar] [CrossRef]
  45. Liu, Z.; Cui, C.; Xu, P.; Dang, R.; Cai, H.; Liao, D.; Yang, M.; Feng, Q.; Yan, X.; Jiang, P. Curcumin Activates AMPK Pathway and Regulates Lipid Metabolism in Rats Following Prolonged Clozapine Exposure. Front. Neurosci. 2017, 11, 558. [Google Scholar] [CrossRef] [Green Version]
  46. Cao, J.; Chen, H.; Lu, W.; Wu, Y.; Wu, X.; Xia, D.; Zhu, J. Myricetin Induces Protective Autophagy by Inhibiting the Phosphorylation of mTOR in HepG2 Cells. Anat. Rec. 2018, 301, 786–795. [Google Scholar] [CrossRef]
  47. Akindehin, S.; Jung, Y.S.; Kim, S.N.; Son, Y.H.; Lee, I.; Seong, J.K.; Jeong, H.W.; Lee, Y.H. Myricetin exerts anti-obesity effects through upregulation of SIRT3 in adipose tissue. Nutrients 2018, 10, 1962. [Google Scholar] [CrossRef]
  48. Jung, H.Y.; Lee, D.; Ryu, H.G.; Choi, B.H.; Go, Y.; Lee, N.; Lee, D.; Son, H.G.; Jeon, J.; Kim, S.H.; et al. Myricetin improves endurance capacity and mitochondrial density by activating SIRT1 and PGC-1α. Sci. Rep. 2017, 7, 6237. [Google Scholar] [CrossRef]
  49. Ryu, D.; Mouchiroud, L.; Andreux, P.A.; Katsyuba, E.; Moullan, N.; Nicolet-Dit-Félix, A.A.; Williams, E.G.; Jha, P.; Lo Sasso, G.; Huzard, D.; et al. Urolithin A induces mitophagy and prolongs lifespan in C. elegans and increases muscle function in rodents. Nat. Med. 2016, 22, 879–888. [Google Scholar] [CrossRef] [PubMed]
  50. Gong, Z.; Huang, J.; Xu, B.; Ou, Z.; Zhang, L.; Lin, X.; Ye, X.; Kong, X.; Long, D.; Sun, X.; et al. Urolithin A attenuates memory impairment and neuroinflammation in APP/PS1 mice. J. Neuroinflammation 2019, 16, 62. [Google Scholar] [CrossRef] [PubMed]
  51. Ono, K.; Hirohata, M.; Yamada, M. Ferulic acid destabilizes preformed β-amyloid fibrils in vitro. Biochem. Biophys. Res. Commun. 2005, 336, 444–449. [Google Scholar] [CrossRef] [PubMed]
  52. Maurya, D.K.; Devasagayam, T.P.A. Antioxidant and prooxidant nature of hydroxycinnamic acid derivatives ferulic and caffeic acids. Food Chem. Toxicol. 2010, 48, 3369–3373. [Google Scholar] [CrossRef] [PubMed]
  53. Yan, J.J.; Jung, J.S.; Kim, T.K.; Hasan, M.A.; Hong, C.W.; Nam, J.S.; Song, D.K. Protective effects of ferulic acid in amyloid precursor protein plus presenilin-1 transgenic mouse model of Alzheimer disease. Biol. Pharm. Bull. 2013, 36, 140–143. [Google Scholar] [CrossRef] [PubMed]
  54. Bian, Z.; Furuya, N.; Zheng, D.M.; Trejo, J.A.O.; Tada, N.; Ezaki, J.; Ueno, T. Ferulic acid induces mammalian target of rapamycin inactivation in cultured mammalian cells. Biol. Pharm. Bull. 2013, 36, 120–124. [Google Scholar] [CrossRef]
  55. Chen, Y.; Zheng, R.; Jia, Z.; Ju, Y. Flavonoids as superoxide scavengers and antioxidants. Free Radic. Biol. Med. 1990, 9, 19–21. [Google Scholar] [CrossRef]
  56. Cho, H.-I.; Park, J.-H.; Choi, H.-S.; Kwak, J.H.; Lee, D.-U.; Lee, S.K.; Lee, S.-M. Protective Mechanisms of Acacetin against d-Galactosamine and Lipopolysaccharide-Induced Fulminant Hepatic Failure in Mice. J. Nat. Prod. 2014, 77, 2497–2503. [Google Scholar] [CrossRef]
  57. Wang, X.; Perumalsamy, H.; Kwon, H.W.; Na, Y.E.; Ahn, Y.J. Effects and possible mechanisms of action of acacetin on the behavior and eye morphology of Drosophila models of Alzheimer’s disease. Sci. Rep. 2015, 5, 16127. [Google Scholar] [CrossRef]
  58. Chang, W.; Wu, Q.Q.; Xiao, Y.; Jiang, X.H.; Yuan, Y.; Zeng, X.F.; Tang, Q.Z. Acacetin protects against cardiac remodeling after myocardial infarction by mediating MAPK and PI3K/Akt signal pathway. J. Pharmacol. Sci. 2017, 135, 156–163. [Google Scholar] [CrossRef]
  59. Li, Y.; Zhao, J.; Holscher, C. Therapeutic Potential of Baicalein in Alzheimer’s Disease and Parkinson’s Disease. CNS Drugs 2017, 31, 639–652. [Google Scholar] [CrossRef]
  60. Lee, H.J.; Noh, Y.H.; Lee, D.Y.; Kim, Y.S.; Kim, K.Y.; Chung, Y.H.; Lee, W.B.; Kim, S.S. Baicalein attenuates 6-hydroxydopamine-induced neurotoxicity in SH-SY5Y cells. Eur. J. Cell Biol. 2005, 84, 897–905. [Google Scholar] [CrossRef]
  61. Zhang, S.Q.; Obregon, D.; Ehrhart, J.; Deng, J.; Tian, J.; Hou, H.; Giunta, B.; Sawmiller, D.; Tan, J. Baicalein reduces β-amyloid and promotes nonamyloidogenic amyloid precursor protein processing in an Alzheimer’s disease transgenic mouse model. J. Neurosci. Res. 2013, 91, 1239–1246. [Google Scholar] [CrossRef] [PubMed]
  62. Lu, J.H.; Ardah, M.T.; Durairajan, S.S.K.; Liu, L.F.; Xie, L.X.; Fong, W.F.D.; Hasan, M.Y.; Huang, J.D.; El-Agnaf, O.M.A.; Li, M. Baicalein Inhibits Formation of α-Synuclein Oligomers within Living Cells and Prevents Aβ Peptide Fibrillation and Oligomerisation. ChemBioChem 2011, 12, 615–624. [Google Scholar] [CrossRef] [PubMed]
  63. Li, J.; Ma, J.; Wang, K.S.; Mi, C.; Wang, Z.; Piao, L.X.; Xu, G.H.; Li, X.; Lee, J.J.; Jin, X. Baicalein inhibits TNF-α-induced NF-κB activation and expression of NF-κB-regulated target gene products. Oncol. Rep. 2016, 36, 2771–2776. [Google Scholar] [CrossRef] [PubMed]
  64. Liu, C.; Wu, J.; Xu, K.; Cai, F.; Gu, J.; Ma, L.; Chen, J. Neuroprotection by baicalein in ischemic brain injury involves PTEN/AKT pathway. J. Neurochem. 2010, 112, 1500–1512. [Google Scholar] [CrossRef]
  65. Gu, X.H.; Xu, L.J.; Liu, Z.Q.; Wei, B.; Yang, Y.J.; Xu, G.G.; Yin, X.P.; Wang, W. The flavonoid baicalein rescues synaptic plasticity and memory deficits in a mouse model of Alzheimer’s disease. Behav. Brain Res. 2016, 311, 309–321. [Google Scholar] [CrossRef]
  66. Angeloni, C.; Barbalace, M.C.; Hrelia, S. Icariin and Its Metabolites as Potential Protective Phytochemicals Against Alzheimer’s Disease. Front. Pharmacol. 2019, 10, 271. [Google Scholar] [CrossRef]
  67. Luo, Y.; Nie, J.; Gong, Q.H.; Lu, Y.F.; Wu, Q.; Shi, J.S. Protective effects of icariin against learning and memory deficits induced by aluminium in rats. Clin. Exp. Pharmacol. Physiol. 2007, 34, 792–795. [Google Scholar] [CrossRef]
  68. Wang, L.; Zhang, L.; Chen, Z.B.; Wu, J.Y.; Zhang, X.; Xu, Y. Icariin enhances neuronal survival after oxygen and glucose deprivation by increasing SIRT1. Eur. J. Pharmacol. 2009, 609, 40–44. [Google Scholar] [CrossRef]
  69. Li, W.W.; Gao, X.M.; Wang, X.M.; Guo, H.; Zhang, B.L. Icariin inhibits hydrogen peroxide-induced toxicity through inhibition of phosphorylation of JNK/p38 MAPK and p53 activity. Mutat. Res. Fundam. Mol. Mech. Mutagenesis 2011, 708, 1–10. [Google Scholar] [CrossRef]
  70. Shi, D.B.; Li, X.X.; Zheng, H.T.; Li, D.W.; Cai, G.X.; Peng, J.J.; Gu, W.L.; Guan, Z.Q.; Xu, Y.; Cai, S.J. Icariin-Mediated Inhibition of NF-κB Activity Enhances the In Vitro and In Vivo Antitumour Effect of 5-Fluorouracil in Colorectal Cancer. Cell Biochem. Biophys. 2014, 69, 523–530. [Google Scholar] [CrossRef]
  71. Li, F.; Dong, H.X.; Gong, Q.H.; Wu, Q.; Jin, F.; Shi, J.S. Icariin decreases both APP and Aβ levels and increases neurogenesis in the brain of Tg2576 mice. Neuroscience 2015, 304, 29–35. [Google Scholar] [CrossRef] [PubMed]
  72. Song, Y.X.; Miao, J.Y.; Qiang, M.; He, R.Q.; Wang, X.M.; Li, W.W. Icariin protects SH-SY5Y cells from formaldehyde-induced injury through suppression of Tau phosphorylation. Chin. J. Integr. Med. 2016, 22, 430–437. [Google Scholar] [CrossRef] [PubMed]
  73. Sheng, C.; Xu, P.; Zhou, K.; Deng, D.; Zhang, C.; Wang, Z. Icariin Attenuates Synaptic and Cognitive Deficits in an Aβ1–42-Induced Rat Model of Alzheimer’s Disease. Biomed. Res. Int. 2017, 2017, 7464872. [Google Scholar] [CrossRef] [PubMed]
  74. Nakajima, A.; Ohizumi, Y. Potential benefits of nobiletin, a citrus flavonoid, against Alzheimer’s disease and Parkinson’s disease. Int. J. Mol. Sci. 2019, 20, 3380. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, L.; Zhao, H.; Zhang, X.; Chen, L.; Zhao, X.; Bai, X.; Zhang, J. Nobiletin protects against cerebral ischemia via activating the p-Akt, p-CREB, BDNF and Bcl-2 pathway and ameliorating BBB permeability in rat. Brain Res. Bull. 2013, 96, 45–53. [Google Scholar] [CrossRef] [PubMed]
  76. Nakajima, A.; Aoyama, Y.; Shin, E.J.; Nam, Y.; Kim, H.C.; Nagai, T.; Yokosuka, A.; Mimaki, Y.; Yokoi, T.; Ohizumi, Y.; et al. Nobiletin, a citrus flavonoid, improves cognitive impairment and reduces soluble Aβ levels in a triple transgenic mouse model of Alzheimer’s disease (3XTg-AD). Behav. Brain Res. 2015, 289, 69–77. [Google Scholar] [CrossRef]
  77. Zhang, L.; Zhang, X.; Zhang, C.; Bai, X.; Zhang, J.; Zhao, X.; Chen, L.; Wang, L.; Zhu, C.; Cui, L.; et al. Nobiletin promotes antioxidant and anti-inflammatory responses and elicits protection against ischemic stroke in vivo. Brain Res. 2016, 1636, 130–141. [Google Scholar] [CrossRef]
  78. Pierzynowska, K.; Podlacha, M.; Gaffke, L.; Majkutewicz, I.; Mantej, J.; Węgrzyn, A.; Osiadły, M.; Myślińska, D.; Węgrzyn, G. Autophagy-dependent mechanism of genistein-mediated elimination of behavioral and biochemical defects in the rat model of sporadic Alzheimer’s disease. Neuropharmacology 2019, 148, 332–346. [Google Scholar] [CrossRef]
  79. Kazi, A.; Daniel, K.G.; Smith, D.M.; Kumar, N.B.; Dou, Q.P. Inhibition of the proteasome activity, a novel mechanism associated with the tumor cell apoptosis-inducing ability of genistein. Biochem. Pharmacol. 2003, 66, 965–976. [Google Scholar] [CrossRef]
  80. Borrás, C.; Gambini, J.; Gómez-Cabrera, M.C.; Sastre, J.; Pallardó, F.V.; Mann, G.E.; Viña, J. Genistein, a soy isoflavone, up-regulates expression of antioxidant genes: Involvement of estrogen receptors, ERK1/2, and NFκB. FASEB J. 2006, 20, E1476–E1481. [Google Scholar] [CrossRef]
  81. Moskot, M.; Montefusco, S.; Jakóbkiewicz-Banecka, J.; Mozolewski, P.; Wȩgrzyn, A.; Di Bernardo, D.; Wȩgrzyn, G.; Medina, D.L.; Ballabio, A.; Gabig-Cimińska, M. The phytoestrogen genistein modulates lysosomal metabolism and Transcription Factor EB (TFEB) activation. J. Biol. Chem. 2014, 289, 17054–17069. [Google Scholar] [CrossRef] [PubMed]
  82. Kwon, Y. Luteolin as a potential preventive and therapeutic candidate for Alzheimer’s disease. Exp. Gerontol. 2017, 95, 39–43. [Google Scholar] [CrossRef] [PubMed]
  83. Liao, Y.; Xu, Y.; Cao, M.; Huan, Y.; Zhu, L.; Jiang, Y.; Shen, W.; Zhu, G. Luteolin Induces Apoptosis and Autophagy in Mouse Macrophage ANA-1 Cells via the Bcl-2 Pathway. J. Immunol. Res. 2018, 2018, 4623919. [Google Scholar] [CrossRef] [PubMed]
  84. Fang, F.; Li, D.; Pan, H.; Chen, D.; Qi, L.; Zhang, R.; Sun, H. Luteolin inhibits apoptosis and improves cardiomyocyte contractile function through the PI3K/Akt pathway in simulated ischemia/reperfusion. Pharmacology 2011, 88, 149–158. [Google Scholar] [CrossRef] [PubMed]
  85. Nunes, C.; Almeida, L.; Barbosa, R.M.; Laranjinha, J. Luteolin suppresses the JAK/STAT pathway in a cellular model of intestinal inflammation. Food Funct. 2017, 8, 387–396. [Google Scholar] [CrossRef] [PubMed]
  86. Cao, Z.; Zhang, H.; Cai, X.; Fang, W.; Chai, D.; Wen, Y.; Chen, H.; Chu, F.; Zhang, Y. Luteolin Promotes Cell Apoptosis by Inducing Autophagy in Hepatocellular Carcinoma. Cell. Physiol. Biochem. 2018, 43, 1803–1812. [Google Scholar] [CrossRef]
  87. Feng, S.T.; Wang, Z.Z.; Yuan, Y.H.; Sun, H.M.; Chen, N.H.; Zhang, Y. Mangiferin: A multipotent natural product preventing neurodegeneration in Alzheimer’s and Parkinson’s disease models. Pharm. Res. 2019, 146, 104336. [Google Scholar] [CrossRef]
  88. Andreu, G.P.; Delgado, R.; Velho, J.A.; Curti, C.; Vercesi, A.E. Iron complexing activity of mangiferin, a naturally occurring glucosylxanthone, inhibits mitochondrial lipid peroxidation induced by Fe2+-citrate. Eur. J. Pharmacol. 2005, 513, 47–55. [Google Scholar] [CrossRef]
  89. Das, J.; Ghosh, J.; Roy, A.; Sil, P.C. Mangiferin exerts hepatoprotective activity against D-galactosamine induced acute toxicity and oxidative/nitrosative stress via Nrf2-NFκB pathways. Toxicol. Appl. Pharmacol. 2012, 260, 35–47. [Google Scholar] [CrossRef]
  90. Kasbe, P.; Jangra, A.; Lahkar, M. Mangiferin ameliorates aluminium chloride-induced cognitive dysfunction via alleviation of hippocampal oxido-nitrosative stress, proinflammatory cytokines and acetylcholinesterase level. J. Trace Elem. Med. Biol. 2015, 31, 107–112. [Google Scholar] [CrossRef]
  91. Jung, J.-S.; Jung, K.; Kim, D.-H.; Kim, H.-S. Selective inhibition of MMP-9 gene expression by mangiferin in PMA-stimulated human astroglioma cells: Involvement of PI3K/Akt and MAPK signaling pathways. Pharmacol. Res. 2012, 66, 95–103. [Google Scholar] [CrossRef] [PubMed]
  92. Wanka, L.; Iqbal, K.; Schreiner, P.R. The lipophilic bullet hits the targets: Medicinal chemistry of adamantane derivatives. Chem. Rev. 2013, 113, 3516–3604. [Google Scholar] [CrossRef] [PubMed]
  93. Shrivastava, S.K.; Sinha, S.K.; Srivastava, P.; Tripathi, P.N.; Sharma, P.; Tripathi, M.K.; Tripathi, A.; Choubey, P.K.; Waiker, D.K.; Aggarwal, L.M.; et al. Design and development of novel p-aminobenzoic acid derivatives as potential cholinesterase inhibitors for the treatment of Alzheimer’s disease. Bioorg. Chem. 2019, 82, 211–223. [Google Scholar] [CrossRef] [PubMed]
  94. Talesa, V.N. Acetylcholinesterase in Alzheimer’s disease. Mech. Ageing Dev. 2001, 122, 1961–1969. [Google Scholar] [CrossRef]
  95. Mehta, M.; Adem, A.; Sabbagh, M. New acetylcholinesterase inhibitors for Alzheimer’s disease. Int. J. Alzheimer’s Dis. 2012, 2012, 728983. [Google Scholar] [CrossRef]
  96. Arundine, M.; Tymianski, M. Molecular mechanisms of calcium-dependent neurodegeneration in excitotoxicity. Cell Calcium 2003, 34, 325–337. [Google Scholar] [CrossRef]
  97. O’Brien, R.J.; Wong, P.C. Amyloid precursor protein processing and Alzheimer’s disease. Annu. Rev. Neurosci. 2011, 34, 185–204. [Google Scholar] [CrossRef]
  98. Thinakaran, G.; Koo, E.H. Amyloid precursor protein trafficking, processing, and function. J. Biol. Chem. 2008, 283, 29615–29619. [Google Scholar] [CrossRef]
  99. Morgan, C.; Colombres, M.; Nuñez, M.T.; Inestrosa, N.C. Structure and function of amyloid in Alzheimer’s disease. Prog. Neurobiol. 2004, 74, 323–349. [Google Scholar] [CrossRef]
  100. Kojro, E.; Fahrenholz, F. The non-amyloidogenic pathway: Structure and function of alpha-secretases. Sub-Cell. Biochem. 2005, 38, 105–127. [Google Scholar]
  101. Menting, K.W.; Claassen, J.A.H.R. β-secretase inhibitor; a promising novel therapeutic drug in Alzheimer’s Disease. Front. Aging Neurosci. 2014, 6, 1–20. [Google Scholar] [CrossRef] [PubMed]
  102. Cole, S.L.; Vassar, R. The Alzheimer’s disease β-secretase enzyme, BACE1. Mol. Neurodegener. 2007, 2, 22. [Google Scholar] [CrossRef] [PubMed]
  103. Farzan, M.; Schnitzler, C.E.; Vasilieva, N.; Leung, D.; Choe, H. BACE2, a β-secretase homolog, cleaves at the β site and within the amyloid-β region of the amyloid-β precursor protein. Proc. Natl. Acad. Sci. USA 2000, 97, 9712–9717. [Google Scholar] [CrossRef] [PubMed]
  104. Cai, H.; Wang, Y.; McCarthy, D.; Wen, H.; Borchelt, D.R.; Price, D.L.; Wong, P.C. BACE1 is the major β-secretase for generation of Aβ peptides by neurons. Nat. Neurosci. 2001, 4, 233–234. [Google Scholar] [CrossRef]
  105. Yu, N.; Hayik, S.A.; Wang, B.; Liao, N.; Reynolds, C.H.; Merz, K.M., Jr. Assigning the protonation states of the key aspartates in β-secretase using QM/MM X-ray structure refinement. J. Chem. Theory Comput. 2006, 2, 1057–1069. [Google Scholar] [CrossRef]
  106. Hernández-Rodríguez, M.; Correa-Basurto, J.; Gutiérrez, A.; Vitorica, J.; Rosales-Hernández, M.C. Asp32 and Asp228 determine the selective inhibition of BACE1 as shown by docking and molecular dynamics simulations. Eur. J. Med. Chem. 2016, 124, 1142–1154. [Google Scholar] [CrossRef]
  107. Sabbah, D.A.; Zhong, H.A. Modeling the protonation states of β-secretase binding pocket by molecular dynamics simulations and docking studies. J. Mol. Graph. Model. 2016, 68, 206–215. [Google Scholar] [CrossRef]
  108. Hong, L.; Turner Iii, R.T.; Koelsch, G.; Shin, D.; Ghosh, A.K.; Tang, J. Crystal structure of memapsin 2 (β-secretase) in complex with an inhibitor OM00-3. Biochemistry 2002, 41, 10963–10967. [Google Scholar] [CrossRef]
  109. Shuto, D.; Kasai, S.; Kimura, T.; Liu, P.; Hidaka, K.; Hamada, T.; Shibakawa, S.; Hayashi, Y.; Hattori, C.; Szabo, B.; et al. KMI-008, a novel β-Secretase inhibitor containing a hydroxymethylcarbonyl isostere as a transition-State mimic: Design and synthesis of substrate-based octapeptides. Bioorg. Med. Chem. Lett. 2003, 13, 4273–4276. [Google Scholar] [CrossRef]
  110. Dineen, T.A.; Weiss, M.M.; Williamson, T.; Acton, P.; Babu-Khan, S.; Bartberger, M.D.; Brown, J.; Chen, K.; Cheng, Y.; Citron, M.; et al. Design and synthesis of potent, orally efficacious hydroxyethylamine derived β-site amyloid precursor protein cleaving enzyme (BACE1) inhibitors. J. Med. Chem. 2012, 55, 9025–9044. [Google Scholar] [CrossRef]
  111. Huang, W.H.; Sheng, R.; Hu, Y.Z. Progress in the development of nonpeptidomimetic BACE 1 inhibitors for Alzheimer’s disease. Curr. Med. Chem. 2009, 16, 1806–1820. [Google Scholar] [CrossRef] [PubMed]
  112. Vega-Hissi, E.G.; Tosso, R.; Enriz, R.D.; Gutierrez, L.J. Molecular insight into the interaction mechanisms of amino-2H-imidazole derivatives with BACE1 protease: A QM/MM and QTAIM study. Int. J. Quantum Chem. 2015, 115, 389–397. [Google Scholar] [CrossRef]
  113. Hansen, M.M.; Jarmer, D.J.; Arslantas, E.; DeBaillie, A.C.; Frederick, A.L.; Harding, M.; Hoard, D.W.; Hollister, A.; Huber, D.; Kolis, S.P.; et al. Synthesis of BACE Inhibitor LY2886721. Part II. Isoxazolidines as Precursors to Chiral Aminothiazines, Selective Peptide Coupling, and a Controlled Reactive Crystallization. Org. Process Res. Dev. 2015, 19, 1214–1230. [Google Scholar] [CrossRef]
  114. Stauffer, S.R.; Graham, S.L. Bicyclic Spiropiperidine Beta-Secretase Inhibitors for the Treatment of Alzheimer’s Disease. U.S. Patent No 8,377,954, 2013. [Google Scholar]
  115. Van Es, J.H.; Van Gijn, M.E.; Riccio, O.; Van Den Born, M.; Vooijs, M.; Begthel, H.; Cozijnsen, M.; Robine, S.; Winton, D.J.; Radtke, F.; et al. Notch/γ-secretase inhibition turns proliferative cells in intestinal crypts and adenomas into goblet cells. Nature 2005, 435, 959–963. [Google Scholar] [CrossRef] [PubMed]
  116. Grill, J.D.; Cummings, J.L. Current therapeutic targets for the treatment of Alzheimer’s disease. Expert Rev. Neurother. 2010, 10, 711–728. [Google Scholar] [CrossRef] [PubMed]
  117. Mayer, S.C.; Kreft, A.F.; Harrison, B.; Abou-Gharbia, M.; Antane, M.; Aschmies, S.; Atchison, K.; Chlenov, M.; Cole, D.C.; Comery, T.; et al. Discovery of begacestat, a Notch-1-sparing γ-secretase inhibitor for the treatment of Alzheimer’s disease. J. Med. Chem. 2008, 51, 7348–7351. [Google Scholar] [CrossRef]
  118. Brahmachari, S.; Paul, A.; Segal, D.; Gazit, E. Inhibition of amyloid oligomerization into different supramolecular architectures by small molecules: Mechanistic insights and design rules. Future Med. Chem. 2017, 9, 797–810. [Google Scholar] [CrossRef]
  119. Aisen, P.A.; Mehran, M.; Poole, R. Clinical data on Alzhemed after 12 months of treatment in patients with mild to moderate Alzheimer’s disease. Neurobiol Aging 2004, 25, S20. [Google Scholar] [CrossRef]
  120. Coman, H.; Nemeş, B. New Therapeutic Targets in Alzheimer’s Disease. Int. J. Gerontol. 2017, 11, 2–6. [Google Scholar] [CrossRef]
  121. Lannfelt, L.; Blennow, K.; Zetterberg, H.; Batsman, S.; Ames, D.; Harrison, J.; Masters, C.L.; Targum, S.; Bush, A.I.; Murdoch, R.; et al. Safety, efficacy, and biomarker findings of PBT2 in targeting Aβ as a modifying therapy for Alzheimer’s disease: A phase IIa, double-blind, randomised, placebo-controlled trial. Lancet Neurol. 2008, 7, 779–786. [Google Scholar] [CrossRef]
  122. Wang, J.; Ho, L.; Zhao, W.; Ono, K.; Rosensweig, C.; Chen, L.; Humala, N.; Teplow, D.B.; Pasinetti, G.M. Grape-derived polyphenolics prevent Aβ oligomerization and attenuate cognitive deterioration in a mouse model of Alzheimer’s disease. J. Neurosci. 2008, 28, 6388–6392. [Google Scholar] [CrossRef] [PubMed]
  123. Cai, Z.; Liu, N.; Wang, C.; Qin, B.; Zhou, Y.; Xiao, M.; Chang, L.; Yan, L.J.; Zhao, B. Role of RAGE in Alzheimer’s Disease. Cell. Mol. Neurobiol. 2016, 36, 483–495. [Google Scholar] [CrossRef] [PubMed]
  124. Orgogozo, J.M.; Gilman, S.; Dartigues, J.F.; Laurent, B.; Puel, M.; Kirby, L.C.; Jouanny, P.; Dubois, B.; Eisner, L.; Flitman, S.; et al. Subacute meningoencephalitis in a subset of patients with AD after Aβ42 immunization. Neurology 2003, 61, 46–54. [Google Scholar] [CrossRef] [PubMed]
  125. Joseph-Mathurin, N.; Dorieux, O.; Trouche, S.G.; Boutajangout, A.; Kraska, A.; Fontès, P.; Verdier, J.M.; Sigurdsson, E.M.; Mestre-Francés, N.; Dhenain, M. Aβ immunization worsens iron deposits in the choroid plexus and cerebral microbleeds. Neurobiol. Aging 2013, 34, 2613–2622. [Google Scholar] [CrossRef]
  126. Friedhoff, P.; Schneider, A.; Mandelkow, E.M.; Mandelkow, E. Rapid assembly of Alzheimer-like paired helical filaments from microtubule-associated protein tau monitored by fluorescence in solution. Biochemistry 1998, 37, 10223–10230. [Google Scholar] [CrossRef]
  127. Kadavath, H.; Hofele, R.V.; Biernat, J.; Kumar, S.; Tepper, K.; Urlaub, H.; Mandelkow, E.; Zweckstetter, M. Tau stabilizes microtubules by binding at the interface between tubulin heterodimers. Proc. Natl. Acad. Sci. USA 2015, 112, 7501–7506. [Google Scholar] [CrossRef] [Green Version]
  128. Ittner, L.M.; Götz, J. Amyloid-β and tau - a toxic pas de deux in Alzheimer’s disease. Nat. Rev. Neurosci. 2011, 12, 67–72. [Google Scholar] [CrossRef]
  129. Šimić, G.; Babić Leko, M.; Wray, S.; Harrington, C.; Delalle, I.; Jovanov-Milošević, N.; Bažadona, D.; Buée, L.; de Silva, R.; Giovanni, G.D.; et al. Tau protein hyperphosphorylation and aggregation in Alzheimer’s disease and other tauopathies, and possible neuroprotective strategies. Biomolecules 2016, 6, 6. [Google Scholar] [CrossRef]
  130. Iqbal, K.; Liu, F.; Gong, C.X. Tau and neurodegenerative disease: The story so far. Nat. Rev. Neurol. 2016, 12, 15–27. [Google Scholar] [CrossRef]
  131. Wang, W.Y.; Tan, M.S.; Yu, J.T.; Tan, L. Role of pro-inflammatory cytokines released from microglia in Alzheimer’s disease. Ann. Transl. Med. 2015, 3, 10. [Google Scholar]
  132. Liu, S.L.; Wang, C.; Jiang, T.; Tan, L.; Xing, A.; Yu, J.T. The Role of Cdk5 in Alzheimer’s Disease. Mol. Neurobiol. 2016, 53, 4328–4342. [Google Scholar] [CrossRef] [PubMed]
  133. Noble, W.; Olm, V.; Takata, K.; Casey, E.; Mary, O.; Meyerson, J.; Gaynor, K.; LaFrancois, J.; Wang, L.; Kondo, T.; et al. Cdk5 is a key factor in tau aggregation and tangle formation in vivo. Neuron 2003, 38, 555–565. [Google Scholar] [CrossRef]
  134. Shupp, A.; Casimiro, M.C.; Pestell, R.G. Biological functions of CDK5 and potential CDK5 targeted clinical treatments. Oncotarget 2017, 8, 17373–17382. [Google Scholar] [CrossRef] [PubMed]
  135. Benson, C.; White, J.; De Bono, J.; O’Donnell, A.; Raynaud, F.; Cruickshank, C.; McGrath, H.; Walton, M.; Workman, P.; Kaye, S.; et al. A phase I trial of the selective oral cyclin-dependent kinase inhibitor seliciclib (CYC202; R-Roscovitine), administered twice daily for 7 days every 21 days. Br. J. Cancer 2007, 96, 29–37. [Google Scholar] [CrossRef] [PubMed]
  136. Mora, A.; Sabio, G.; González-Polo, R.A.; Cuenda, A.; Alessi, D.R.; Alonso, J.C.; Fuentes, J.M.; Soler, G.; Centeno, F. Lithium inhibits caspase 3 activation and dephosphorylation of PKB and GSK3 induced by K+ deprivation in cerebellar granule cells. J. Neurochem. 2001, 78, 199–206. [Google Scholar] [CrossRef]
  137. Martinez, A.; Castro, A.; Dorronsoro, I.; Alonso, M. Glycogen synthase kinase 3 (GSK-3) inhibitors as new promising drugs for diabetes, neurodegeneration, cancer, and inflammation. Med. Res. Rev. 2002, 22, 373–384. [Google Scholar] [CrossRef]
  138. Meijer, L.; Skaltsounis, A.L.; Magiatis, P.; Polychronopoulos, P.; Knockaert, M.; Leost, M.; Ryan, X.P.; Vonica, C.A.; Brivanlou, A.; Dajani, R.; et al. GSK-3-Selective Inhibitors Derived from Tyrian Purple Indirubins. Chem. Biol. 2003, 10, 1255–1266. [Google Scholar] [CrossRef] [Green Version]
  139. King, M.K.; Pardo, M.; Cheng, Y.; Downey, K.; Jope, R.S.; Beurel, E. Glycogen synthase kinase-3 inhibitors: Rescuers of cognitive impairments. Pharmacol. Ther. 2014, 141, 1–12. [Google Scholar] [CrossRef]
  140. Wan, Y.; Hur, W.; Cho, C.Y.; Liu, Y.; Adrian, F.J.; Lozach, O.; Bach, S.; Mayer, T.; Fabbro, D.; Meijer, L.; et al. Synthesis and target identification of hymenialdisine analogs. Chem. Biol. 2004, 11, 247–259. [Google Scholar] [CrossRef]
  141. Kontsekova, E.; Zilka, N.; Kovacech, B.; Novak, P.; Novak, M. First-in-man tau vaccine targeting structural determinants essential for pathological tau-tau interaction reduces tau oligomerisation and neurofibrillary degeneration in an Alzheimer’s disease model. Alzheimer’s Res. Ther. 2014, 6, 44. [Google Scholar] [CrossRef]
  142. Novak, P.; Schmidt, R.; Kontsekova, E.; Zilka, N.; Kovacech, B.; Skrabana, R.; Vince-Kazmerova, Z.; Katina, S.; Fialova, L.; Prcina, M.; et al. Safety and immunogenicity of the tau vaccine AADvac1 in patients with Alzheimer’s disease: A randomised, double-blind, placebo-controlled, phase 1 trial. Lancet Neurol. 2017, 16, 123–134. [Google Scholar] [CrossRef]
  143. Boyles, J.K.; Zoellner, C.D.; Anderson, L.J.; Kosik, L.M.; Pitas, R.E.; Weisgraber, K.H.; Hui, D.Y.; Mahley, R.W.; Gebicke-Haerter, P.J.; Ignatius, M.J.; et al. A role for apolipoprotein E, apolipoprotein A-I, and low density lipoprotein receptors in cholesterol transport during regeneration and remyelination of the rat sciatic nerve. J. Clin. Investig. 1989, 83, 1015–1031. [Google Scholar] [CrossRef] [PubMed]
  144. Michaelson, D.M. APOE ε4: The most prevalent yet understudied risk factor for Alzheimer’s disease. Alzheimer Dement. 2014, 10, 861–868. [Google Scholar] [CrossRef]
  145. Wisniewski, T.; Golabek, A.; Matsubara, E.; Ghiso, J.; Frangione, B. Apolipoprotein E: Binding to Soluble Alzheimer′s β-Amyloid. Biochem. Biophys. Res. Commun. 1993, 192, 359–365. [Google Scholar] [CrossRef] [PubMed]
  146. Galpern, W.R.; Lang, A.E. Interface between tauopathies and synucleinopathies: A tale of two proteins. Ann. Neurol. 2006, 59, 449–458. [Google Scholar] [CrossRef]
  147. Hashimoto, M.; Yasuda, M.; Tanimukai, S.; Matsui, M.; Hirono, N.; Kazui, H.; Mori, E. Apolipoprotein E ε4 and the pattern of regional brain atrophy in Alzheimer’s disease. Neurology 2001, 57, 1461–1466. [Google Scholar] [CrossRef]
  148. Giau, V.V.; Bagyinszky, E.; An, S.S.A.; Kim, S.Y. Role of apolipoprotein E in neurodegenerative diseases. Neuropsychiatr. Dis. Treat. 2015, 11, 1723–1737. [Google Scholar] [CrossRef]
  149. Mahley, R.W.; Weisgraber, K.H.; Huang, Y. Apolipoprotein E4: A causative factor and therapeutic target in neuropathology, including Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2006, 103, 5644–5651. [Google Scholar] [CrossRef]
  150. Ungar, L.; Altmann, A.; Greicius, M.D. Apolipoprotein E, gender, and Alzheimer’s disease: An overlooked, but potent and promising interaction. Brain Imaging Behav. 2014, 8, 262–273. [Google Scholar] [CrossRef]
  151. Praticò, D. Oxidative stress hypothesis in Alzheimer’s disease: A reappraisal. Trends Pharmacol. Sci. 2008, 29, 609–615. [Google Scholar] [CrossRef]
  152. Lin, M.T.; Beal, M.F. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 2006, 443, 787–795. [Google Scholar] [CrossRef] [PubMed]
  153. Kurz, T.; Terman, A.; Brunk, U.T. Autophagy, ageing and apoptosis: The role of oxidative stress and lysosomal iron. Arch. Biochem. Biophys. 2007, 462, 220–230. [Google Scholar] [CrossRef] [PubMed]
  154. Chen, X.; Petranovic, D. Amyloid-ß peptide-induced cytotoxicity and mitochondrial dysfunction in yeast. FEMS Yeast Res. 2015, 15, 6. [Google Scholar] [CrossRef] [PubMed]
  155. Rizzi, F.; Trougakos, I.P.; Pintus, G.; Sykiotis, G.P. Redox Status and Proteostasis in Ageing and Disease. Oxidative Med. Cell. Longev. 2016, 2016. [Google Scholar] [CrossRef] [PubMed]
  156. Mittal, M.; Siddiqui, M.R.; Tran, K.; Reddy, S.P.; Malik, A.B. Reactive oxygen species in inflammation and tissue injury. Antioxid. Redox Signal 2014, 20, 1126–1167. [Google Scholar] [CrossRef] [PubMed]
  157. Kandola, K.; Bowman, A.; Birch-Machin, M.A. Oxidative stress—A key emerging impact factor in health, ageing, lifestyle and aesthetics. Int. J. Cosmet. Sci. 2015, 37, 1–8. [Google Scholar] [CrossRef]
  158. Kumar, A.; Singh, A. A review on mitochondrial restorative mechanism of antioxidants in Alzheimer’s disease and other neurological conditions. Front. Pharmacol. 2015, 6, 206. [Google Scholar] [CrossRef]
  159. Caine, J.; Sankovich, S.; Antony, H.; Waddington, L.; Macreadie, P.; Varghese, J.; Macreadie, I. Alzheimer’s Aβ fused to green fluorescent protein induces growth stress and a heat shock response. FEMS Yeast Res. 2007, 7, 1230–1236. [Google Scholar] [CrossRef]
  160. Deibel, M.A.; Ehmann, W.D.; Markesbery, W.R. Copper, iron, and zinc imbalances in severely degenerated brain regions in Alzheimer’s disease: Possible relation to oxidative stress. J. Neurol. Sci. 1996, 143, 137–142. [Google Scholar] [CrossRef]
  161. Tomljenovic, L. Aluminum and Alzheimer’s disease: After a century of controversy, is there a plausible link? J. Alzheimer Dis. 2011, 23, 567–598. [Google Scholar] [CrossRef]
  162. Karlsson, M.; Frennesson, C.; Gustafsson, T.; Brunk, U.T.; Nilsson, S.E.G.; Kurz, T. Autophagy of iron-binding proteins may contribute to the oxidative stress resistance of ARPE-19 cells. Exp. Eye Res. 2013, 116, 359–365. [Google Scholar] [CrossRef] [PubMed]
  163. Zhao, Y.; Zhao, B. Oxidative stress and the pathogenesis of Alzheimer’s disease. Oxidative Med. Cell. Longev. 2013. [Google Scholar] [CrossRef]
  164. Reddy, P.H.; Beal, M.F. Amyloid beta, mitochondrial dysfunction and synaptic damage: Implications for cognitive decline in aging and Alzheimer’s disease. Trends Mol. Med. 2008, 14, 45–53. [Google Scholar] [CrossRef] [PubMed]
  165. Poprac, P.; Jomova, K.; Simunkova, M.; Kollar, V.; Rhodes, C.J.; Valko, M. Targeting Free Radicals in Oxidative Stress-Related Human Diseases. Trends Pharmacol. Sci. 2017, 38, 592–607. [Google Scholar] [CrossRef] [PubMed]
  166. Anekonda, T.S. Resveratrol-a boon for treating Alzheimer’s disease? Brain Res. Rev. 2006, 52, 316–326. [Google Scholar] [CrossRef]
  167. Khan, H.; Marya; Amin, S.; Kamal, M.A.; Patel, S. Flavonoids as acetylcholinesterase inhibitors: Current therapeutic standing and future prospects. Biomed. Pharmacother. 2018, 101, 860–870. [Google Scholar] [CrossRef]
  168. Ding, X.; Ouyang, M.A.; Liu, X.; Wang, R.Z. Acetylcholinesterase inhibitory activities of flavonoids from the leaves of Ginkgo biloba against brown planthopper. J. Chem. 2013, 2013, 645086. [Google Scholar] [CrossRef]
  169. Cox, C.J.; Choudhry, F.; Peacey, E.; Perkinton, M.S.; Richardson, J.C.; Howlett, D.R.; Lichtenthaler, S.F.; Francis, P.T.; Williams, R.J. Dietary (-)-epicatechin as a potent inhibitor of βγ-secretase amyloid precursor protein processing. Neurobiol. Aging 2015, 36, 178–187. [Google Scholar] [CrossRef]
  170. Kook, S.Y.; Lee, K.M.; Kim, Y.; Cha, M.Y.; Kang, S.; Baik, S.H.; Lee, H.; Park, R.; Mook-Jung, I. High-dose of vitamin C supplementation reduces amyloid plaque burden and ameliorates pathological changes in the brain of 5XFAD mice. Cell Death Dis. 2014, 5, e1083. [Google Scholar] [CrossRef]
  171. Zhang, L.F.; Zhou, Z.W.; Wang, Z.H.; Du, Y.H.; He, Z.X.; Cao, C.; Zhou, S.F. Coffee and caffeine potentiate the antiamyloidogenic activity of melatonin via inhibition of aβ oligomerization and modulation of the Tau-mediated pathway in N2a/APP cells. Drug Des. Dev. Ther. 2015, 9, 241–272. [Google Scholar]
  172. Vacek, J.C.; Behera, J.; George, A.K.; Kamat, P.K.; Kalani, A.; Tyagi, N. Tetrahydrocurcumin ameliorates homocysteine-mediated mitochondrial remodeling in brain endothelial cells. J. Cell. Physiol. 2018, 233, 3080–3092. [Google Scholar] [CrossRef] [PubMed]
  173. Mancino, A.M.; Hindo, S.S.; Kochi, A.; Lim, M.H. Effects of Clioquinol on Metal-Triggered Amyloid-β Aggregation Revisited. Inorg. Chem. 2009, 48, 9596–9598. [Google Scholar] [CrossRef] [PubMed]
  174. Gomes, L.M.F.; Vieira, R.P.; Jones, M.R.; Wang, M.C.P.; Dyrager, C.; Souza-Fagundes, E.M.; Da Silva, J.G.; Storr, T.; Beraldo, H. 8-Hydroxyquinoline Schiff-base compounds as antioxidants and modulators of copper-mediated Aβ peptide aggregation. J. Inorg. Biochem. 2014, 139, 106–116. [Google Scholar] [CrossRef] [PubMed]
  175. Liang, S.H.; Southon, A.G.; Fraser, B.H.; Krause-Heuer, A.M.; Zhang, B.; Shoup, T.M.; Lewis, R.; Volitakis, I.; Han, Y.; Greguric, I.; et al. Novel Fluorinated 8-Hydroxyquinoline Based Metal Ionophores for Exploring the Metal Hypothesis of Alzheimer’s Disease. ACS Med. Chem. Lett. 2015, 6, 1025–1029. [Google Scholar] [CrossRef] [PubMed]
  176. Morris, D.R.; Fillingame, R.H. Regulation of amino acid decarboxylation. Annu. Rev. Biochem. 1974, 43, 303–325. [Google Scholar] [CrossRef]
  177. Bai, X.; Edden, R.A.E.; Gao, F.; Wang, G.; Wu, L.; Zhao, B.; Wang, M.; Chan, Q.; Chen, W.; Barker, P.B. Decreased γ-aminobutyric acid levels in the parietal region of patients with Alzheimer’s disease. J. Magn. Reson. Imaging 2015, 41, 1326–1331. [Google Scholar] [CrossRef]
  178. Sternfeld, F.; Carling, R.W.; Jelley, R.A.; Ladduwahetty, T.; Merchant, K.J.; Moore, K.W.; Reeve, A.J.; Street, L.J.; O’Connor, D.; Sohal, B.; et al. Selective, Orally Active γ-Aminobutyric AcidA α5 Receptor Inverse Agonists as Cognition Enhancers. J. Med. Chem. 2004, 47, 2176–2179. [Google Scholar] [CrossRef]
  179. Lee, J.Y.; Friedman, J.E.; Angel, I.; Kozak, A.; Koh, J.Y. The lipophilic metal chelator DP-109 reduces amyloid pathology in brains of human β-amyloid precursor protein transgenic mice. Neurobiol. Aging 2004, 25, 1315–1321. [Google Scholar] [CrossRef]
  180. Rose, G.M.; Hopper, A.; De Vivo, M.; Tehim, A. Phosphodiesterase inhibitors for cognitive enhancement. Curr. Pharm. Des. 2005, 11, 3329–3334. [Google Scholar] [CrossRef]
  181. Li, J.; Liu, C.N.; Wei, N.; Li, X.D.; Liu, Y.Y.; Yang, R.; Jia, Y.J. Protective effects of BAY 73–6691, a selective inhibitor of phosphodiesterase 9, on amyloid-β peptides-induced oxidative stress in in-vivo and in-vitro models of Alzheimer’s disease. Brain Res. 2016, 1642, 327–335. [Google Scholar] [CrossRef]
  182. Hagen, T.J.; Mo, X.; Burgin, A.B.; Fox Iii, D.; Zhang, Z.; Gurney, M.E. Discovery of triazines as selective PDE4B versus PDE4D inhibitors. Bioorg. Med. Chem. Lett. 2014, 24, 4031–4034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Blokland, A.; Schreiber, R.; Prickaerts, J. Improving memory: A role for phosphodiesterases. Curr. Pharm. Des. 2006, 12, 2511–2523. [Google Scholar] [CrossRef] [PubMed]
  184. Qin, W.; Ho, L.; Pompl, P.N.; Peng, Y.; Zhao, Z.; Xiang, Z.; Robakis, N.K.; Shioi, J.; Suh, J.; Pasinetti, G.M. Cyclooxygenase (COX)-2 and COX-1 Potentiate β-Amyloid Peptide Generation through Mechanisms That Involve γ-Secretase Activity. J. Biol. Chem. 2003, 278, 50970–50977. [Google Scholar] [CrossRef] [PubMed]
  185. Hewett, S.J.; Uliasz, T.F.; Vidwans, A.S.; Hewett, J.A. Cyclooxygenase-2 contributes to N-methyl-D-aspartate-mediated neuronal cell death in primary cortical cell culture. J. Pharmacol. Exp. Ther. 2000, 293, 417–425. [Google Scholar] [PubMed]
  186. Park, S.A.; Chevallier, N.; Tejwani, K.; Hung, M.M.; Maruyama, H.; Golde, T.E.; Koo, E.H. Deficiency in either COX-1 or COX-2 genes does not affect amyloid beta protein burden in amyloid precursor protein transgenic mice. Biochem. Biophys. Res. Commun. 2016, 478, 286–292. [Google Scholar] [CrossRef]
  187. Brioni, J.D.; Esbenshade, T.A.; Garrison, T.R.; Bitner, S.R.; Cowart, M.D. Discovery of histamine H3 antagonists for the treatment of cognitive disorders and Alzheimer’s disease. J. Pharmacol. Exp. Ther. 2011, 336, 38–46. [Google Scholar] [CrossRef]
  188. Delay-Goyet, P.; Blanchard, V.; Schussler, N.; Lopez-Grancha, M.; Ménager, J.; Mary, V.; Sultan, E.; Buzy, A.; Guillemot, J.C.; Stemmelin, J.; et al. SAR110894, a potent histamine H3-receptor antagonist, displays disease-modifying activity in a transgenic mouse model of tauopathy. Alzheimer Dement. Transl. Res. Clin. Interv. 2016, 2, 267–280. [Google Scholar] [CrossRef] [Green Version]
  189. Yun, H.M.; Park, K.R.; Kim, E.C.; Kim, S.; Hong, J.T. Serotonin 6 receptor controls Alzheimer’s disease and depression. Oncotarget 2015, 6, 26716–26728. [Google Scholar] [CrossRef]
  190. Claeysen, S.; Bockaert, J.; Giannoni, P. Serotonin: A New Hope in Alzheimer’s Disease? Acs Chem. Neurosci. 2015, 6, 940–943. [Google Scholar] [CrossRef]
  191. Chinetti, G.; Fruchart, J.C.; Staels, B. Peroxisome proliferator-activated receptors (PPARs): Nuclear receptors at the crossroads between lipid metabolism and inflammation. Inflamm. Res. 2000, 49, 497–505. [Google Scholar] [CrossRef]
  192. Braissant, O.; Foufelle, F.; Scotto, C.; Dauça, M.; Wahli, W. Differential expression of peroxisome proliferator-activated receptors (PPARs): Tissue distribution of PPAR-α, -β, and -γ in the adult rat. Endocrinology 1996, 137, 354–366. [Google Scholar] [CrossRef] [PubMed]
  193. Bordet, R.; Ouk, T.; Petrault, O.; Gele, P.; Gautier, S.; Laprais, M.; Deplanque, D.; Duriez, P.; Staels, B.; Fruchart, J. PPAR: A New Pharmacological Target for Neuroprotection in Stroke and Neurodegenerative Diseases. Biochem Soc Trans. 2006, 34, 1341–1346. [Google Scholar] [CrossRef] [PubMed]
  194. Combs, C.K.; Johnson, D.E.; Karlo, J.C.; Cannady, S.B.; Landreth, G.E. Inflammatory mechanisms in Alzheimer’s disease: Inhibition of β- amyloid-stimulated proinflammatory responses and neurotoxicity by PPARγ agonists. J. Neurosci. 2000, 20, 558–567. [Google Scholar] [CrossRef] [PubMed]
  195. Heneka, M.T.; Sastre, M.; Dumitrescu-Ozimek, L.; Hanke, A.; Dewachter, I.; Kuiperi, C.; O’Banion, K.; Klockgether, T.; Van Leuven, F.; Landreth, G.E. Acute treatment with the PPARγ agonist pioglitazone and ibuprofen reduces glial inflammation and Aβ1-42 levels in APPV717I transgenic mice. Brain 2005, 128, 1442–1453. [Google Scholar] [CrossRef]
  196. Cramer, P.E.; Cirrito, J.R.; Wesson, D.W.; Lee, C.Y.D.; Karlo, J.C.; Zinn, A.E.; Casali, B.T.; Restivo, J.L.; Goebel, W.D.; James, M.J.; et al. ApoE-directed therapeutics rapidly clear β-amyloid and reverse deficits in AD mouse models. Science 2012, 335, 1503–1506. [Google Scholar] [CrossRef]
  197. Skerrett, R.; Pellegrino, M.P.; Casali, B.T.; Taraboanta, L.; Landreth, G.E. Combined liver X receptor/peroxisome proliferator-activated receptor γ agonist treatment reduces amyloid β levels and improves behavior in amyloid precursor protein/presenilin 1 mice. J. Biol. Chem. 2015, 290, 21591–21602. [Google Scholar] [CrossRef]
  198. Koster, K.P.; Smith, C.; Valencia-Olvera, A.C.; Thatcher, G.R.J.; LaDu, M.J.; Tai, L.M. Rexinoids as therapeutics for Alzheimer disease: Role of APOE. Curr. Top. Med. Chem. 2016, 16, 708–720. [Google Scholar]
  199. Tong, M.; Dominguez, C.; Didsbury, J.; de la Monte, S.M. Targeting Alzheimer’s disease neuro-metabolic dysfunction with a small molecule nuclear receptor agonist (T3D-959) reverses disease pathologies. J. Alzheimers Dis. Parkinsonism 2016, 6, 238–244. [Google Scholar] [CrossRef]
  200. Watson, G.S.; Cholerton, B.A.; Reger, M.A.; Baker, L.D.; Plymate, S.R.; Asthana, S.; Fishel, M.A.; Kulstad, J.J.; Green, P.S.; Cook, D.G.; et al. Preserved cognition in patients with early Alzheimer disease and amnestic mild cognitive impairment during treatment with rosiglitazone: A preliminary study. J. Am. Assoc. Geriatr. Psychiatry 2005, 13, 950–958. [Google Scholar] [CrossRef]
  201. Janefjord, E.; Mååg, J.L.V.; Harvey, B.S.; Smid, S.D. Cannabinoid effects on β amyloid fibril and aggregate formation, neuronal and microglial-activated neurotoxicity in vitro. Cell. Mol. Neurobiol. 2014, 34, 31–42. [Google Scholar] [CrossRef]
  202. Rampa, A.; Gobbi, S.; Belluti, F.; Bisi, A. Emerging targets in neurodegeneration: New opportunities for Alzheimer’s disease treatment? Curr. Top. Med. Chem. 2013, 13, 1879–1904. [Google Scholar] [CrossRef] [PubMed]
  203. Altamura, C.; Ventriglia, M.; Martini, M.G.; Montesano, D.; Errante, Y.; Piscitelli, F.; Scrascia, F.; Quattrocchi, C.; Palazzo, P.; Seccia, S.; et al. Elevation of plasma 2-arachidonoylglycerol levels in alzheimer’s disease patients as a potential protective mechanism against neurodegenerative decline. J. Alzheimer Dis. 2015, 46, 497–506. [Google Scholar]
  204. Watabiki, T.; Tsuji, N.; Kiso, T.; Ozawa, T.; Narazaki, F.; Kakimoto, S. In vitro and in vivo pharmacological characterization of ASP8477: A novel highly selective fatty acid amide hydrolase inhibitor. Eur. J. Pharmacol. 2017, 815, 42–48. [Google Scholar] [CrossRef] [PubMed]
  205. Bedse, G.; Romano, A.; Lavecchia, A.M.; Cassano, T.; Gaetani, S. The role of endocannabinoid signaling in the molecular mechanisms of neurodegeneration in Alzheimer’s disease. J. Alzheimer Dis. 2014, 43, 1115–1136. [Google Scholar] [CrossRef] [PubMed]
  206. Dias, I.H.K.; Mistry, J.; Fell, S.; Reis, A.; Spickett, C.M.; Polidori, M.C.; Lip, G.Y.H.; Griffiths, H.R. Oxidized LDL lipids increase β-amyloid production by SH-SY5Y cells through glutathione depletion and lipid raft formation. Free Radic. Biol. Med. 2014, 75, 48–59. [Google Scholar] [CrossRef]
  207. Wolozin, B.; Wang, S.W.; Li, N.C.; Lee, A.; Lee, T.A.; Kazis, L.E. Simvastatin is associated with a reduced incidence of dementia and Parkinson’s disease. BMC Med. 2007, 5, 20. [Google Scholar] [CrossRef]
  208. Atta, M. Exploring the relationship between statins and Alzheimer’s disease: Can statins really prevent Alzheimer’s disease? Adv. Alzheimer. Dis. 2015, 4, 10–14. [Google Scholar] [CrossRef]
  209. Geifman, N.; Brinton, R.D.; Kennedy, R.E.; Schneider, L.S.; Butte, A.J. Evidence for benefit of statins to modify cognitive decline and risk in Alzheimer’s disease. Alzheimer Res. Ther. 2017, 9, 10. [Google Scholar] [CrossRef]
  210. Dhakal, S.; Subhan, M.; Fraser, M.J.; Gardiner, K.; Macreadie, I. Simvastatin Efficiently Reduces Levels of Alzheimer’s Amyloid Beta in Yeast. Int. J. Mol. Sci. 2019, 20, 3531. [Google Scholar] [CrossRef]
  211. Terenghi, G. Peripheral nerve regeneration and neurotrophic factors. J Anat 1999, 194, 1–14. [Google Scholar] [CrossRef]
  212. Levy, Y.S.; Gilgun-Sherki, Y.; Melamed, E.; Offen, D. Therapeutic Potential of Neurotrophic Factors in Neurodegenerative Diseases. BioDrugs 2005, 19, 97–127. [Google Scholar] [CrossRef] [PubMed]
  213. Maisonpierre, P.C.; Belluscio, L.; Friedman, B.; Alderson, R.F.; Wiegand, S.J.; Furth, M.E.; Lindsay, R.M.; Yancopoulos, G.D. NT-3, BDNF, and NGF in the developing rat nervous system: Parallel as well as reciprocal patterns of expression. Neuron 1990, 5, 501–509. [Google Scholar] [CrossRef]
  214. Huang, E.J.; Reichardt, L.F. Trk receptors: Roles in neuronal signal transduction. Annu. Rev. Biochem. 2003, 72, 609–642. [Google Scholar] [CrossRef] [PubMed]
  215. Longo, F.M.; Massa, S.M. Small-molecule modulation of neurotrophin receptors: A strategy for the treatment of neurological disease. Nat. Rev. Drug Discov. 2013, 12, 507. [Google Scholar] [CrossRef] [PubMed]
  216. Patapoutian, A.; Reichardt, L.F. Trk receptors: Mediators of neurotrophin action. Curr. Opin. Neurobiol. 2001, 11, 272–280. [Google Scholar] [CrossRef]
  217. Levi-Montalcini, R. The nerve growth factor: Thirty-five years later. EMBO J 1987, 6, 1145–1154. [Google Scholar] [CrossRef] [PubMed]
  218. Korsching, S.; Auburger, G.; Heumann, R.; Scott, J.; Thoenen, H. Levels of nerve growth factor and its mRNA in the central nervous system of the rat correlate with cholinergic innervation. EMBO J 1985, 4, 1389–1393. [Google Scholar] [CrossRef] [PubMed]
  219. Olson, L. NGF and the Treatment of Alzheimer’s Disease. Exp. Neurol. 1993, 124, 5–15. [Google Scholar] [CrossRef]
  220. Salehi, A.; Delcroix, J.D.; Swaab, D. Alzheimer’s disease and NGF signaling. J. Neural Transm. 2004, 111, 323–345. [Google Scholar] [CrossRef]
  221. Capsoni, S.; Cattaneo, A. On the Molecular Basis Linking Nerve Growth Factor (NGF) to Alzheimer’s Disease. Cell. Mol. Neurobiol. 2006, 26, 617–631. [Google Scholar] [CrossRef]
  222. Jakob-Roetne, R.; Jacobsen, H. Alzheimer’s disease: From pathology to therapeutic approaches. Angew. Chem. Int. Ed. 2009, 48, 3030–3059. [Google Scholar] [CrossRef]
  223. Nisticò, R.; Pignatelli, M.; Piccinin, S.; Mercuri, N.B.; Collingridge, G. Targeting synaptic dysfunction in Alzheimer’s disease therapy. Mol. Neurobiol. 2012, 46, 572–587. [Google Scholar] [CrossRef]
  224. Hefti, F. Nerve growth factor promotes survival of septal cholinergic neurons after fimbrial transections. J. Neurosci. 1986, 6, 2155–2162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Fischer, W.; Wictorin, K.; Björklund, A.; Williams, L.; Varon, S.; Gage, F. Amelioration of cholinergic neuron atrophy and spatial memory impairment in aged rats by nerve growth factor. Nature 1987, 329, 65. [Google Scholar] [CrossRef]
  226. Hagg, T.; Manthorpe, M.; Vahlsing, H.L.; Varon, S. Delayed treatment with nerve growth factor reverses the apparent loss of cholinergic neurons after acute brain damage. Exp. Neurol. 1988, 101, 303–312. [Google Scholar] [CrossRef]
  227. Hagg, T.; Vahlsing, H.L.; Manthorpe, M.; Varon, S. Nerve growth factor infusion into the denervated adult rat hippocampal formation promotes its cholinergic reinnervation. J. Neurosci. 1990, 10, 3087–3092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Hagg, T.; Hagg, F.; Vahlsing, H.; Manthorpe, M.; Varon, S. Nerve growth factor effects on cholinergic neurons of neostriatum and nucleus accumbens in the adult rat. Neuroscience 1989, 30, 95–103. [Google Scholar] [CrossRef]
  229. Friden, P.M.; Walus, L.R.; Watson, P.; Doctrow; Kozarich, J.W.; Backman, C.; Bergman, H.; Hoffer, B.; Bloom, F.; Granholm, A.C. Blood-brain barrier penetration and in vivo activity of an NGF conjugate. Science 1993, 259, 373. [Google Scholar] [CrossRef] [PubMed]
  230. Poduslo, J.F.; Curran, G.L. Permeability at the blood-brain and blood-nerve barriers of the neurotrophic factors: NGF, CNTF, NT-3, BDNF. Mol. Brain Res. 1996, 36, 280–286. [Google Scholar] [CrossRef]
  231. Pan, W.; Banks, W.A.; Kastin, A.J. Permeability of the blood–brain barrier to neurotrophins. Brain Res. 1998, 788, 87–94. [Google Scholar] [CrossRef]
  232. Thoenen, H.; Sendtner, M. Neurotrophins: From enthusiastic expectations through sobering experiences to rational therapeutic approaches. Nat. Neurosci. 2002, 5, 1046. [Google Scholar] [CrossRef] [PubMed]
  233. Mitra, S.; Behbahani, H.; Eriksdotter, M. Innovative therapy for Alzheimer’s disease-with focus on biodelivery of NGF. Front. Neurosci. 2019, 13, 38. [Google Scholar] [CrossRef]
  234. Zhu, Y.; Xiao, K.; Ma, L.; Xiong, B.; Fu, Y.; Yu, H.; Wang, W.; Wang, X.; Hu, D.; Peng, H.; et al. Design, synthesis and biological evaluation of novel dual inhibitors of acetylcholinesterase and β-secretase. Bioorg. Med. Chem. 2009, 17, 1600–1613. [Google Scholar] [CrossRef] [PubMed]
  235. Mohamed, T.; Yeung, J.C.K.; Vasefi, M.S.; Beazely, M.A.; Rao, P.P.N. Development and evaluation of multifunctional agents for potential treatment of Alzheimer’s disease: Application to a pyrimidine-2,4-diamine template. Bioorg. Med. Chem. Lett. 2012, 22, 4707–4712. [Google Scholar] [CrossRef] [PubMed]
  236. Fernández-Bachiller, M.I.; Pérez, C.; Monjas, L.; Rademann, J.; Rodríguez-Franco, M.I. New tacrine-4-oxo-4H-chromene hybrids as multifunctional agents for the treatment of Alzheimer’s disease, with cholinergic, antioxidant, and β-amyloid-reducing properties. J. Med. Chem. 2012, 55, 1303–1317. [Google Scholar] [CrossRef] [PubMed]
  237. Camps, P.; El Achab, R.; Morral, J.; Munoz-Torrero, D.; Badia, A.; Eladi Banos, J.; Vivas, N.M.; Barril, X.; Orozco, M.; Javier Luque, F. New tacrine-huperzine A hybrids (huprines): Highly potent tight-binding acetylcholinesterase inhibitors of interest for the treatment of Alzheimer’s Disease. J. Med. Chem. 2000, 43, 4657–4666. [Google Scholar] [CrossRef] [PubMed]
  238. Zha, X.; Lamba, D.; Zhang, L.; Lou, Y.; Xu, C.; Kang, D.; Chen, L.; Xu, Y.; Zhang, L.; De Simone, A.; et al. Novel Tacrine-Benzofuran Hybrids as Potent Multitarget-Directed Ligands for the Treatment of Alzheimers Disease: Design, Synthesis, Biological Evaluation, and X-ray Crystallography. J. Med. Chem. 2016, 59, 114–131. [Google Scholar] [CrossRef] [PubMed]
  239. Gabr, M.T.; Abdel-Raziq, M.S. Design and synthesis of donepezil analogues as dual AChE and BACE-1 inhibitors. Bioorg. Chem. 2018, 80, 245–252. [Google Scholar] [CrossRef]
  240. Zhang, P.; Xu, S.; Zhu, Z.; Xu, J. Multi-target design strategies for the improved treatment of Alzheimer’s disease. Eur. J. Med. Chem. 2019, 228–247. [Google Scholar] [CrossRef]
  241. Prati, F.; De Simone, A.; Bisignano, P.; Armirotti, A.; Summa, M.; Pizzirani, D.; Scarpelli, R.; Perez, D.I.; Andrisano, V.; Perez-Castillo, A.; et al. Multitarget drug discovery for Alzheimer’s disease: Triazinones as BACE-1 and GSK-3β inhibitors. Angew. Chem. Int. Ed. 2015, 54, 1578–1582. [Google Scholar] [CrossRef]
  242. Prasad, S.; Gupta, S.C.; Tyagi, A.K.; Aggarwal, B.B. Curcumin, a component of golden spice: From bedside to bench and back. Biotechnol. Adv. 2014, 32, 1053–1064. [Google Scholar] [CrossRef] [PubMed]
  243. Xie, S.; Chen, J.; Li, X.; Su, T.; Wang, Y.; Wang, Z.; Huang, L.; Li, X. Synthesis and evaluation of selegiline derivatives as monoamine oxidase inhibitor, antioxidant and metal chelator against Alzheimer’s disease. Bioorg. Med. Chem. 2015, 23, 3722–3729. [Google Scholar] [CrossRef] [PubMed]
  244. Wang, Z.; Wang, Y.; Wang, B.; Li, W.; Huang, L.; Li, X. Design, synthesis, and evaluation of orally available clioquinol-moracin M hybrids as multitarget-directed ligands for cognitive improvement in a rat model of neurodegeneration in Alzheimer’s disease. J. Med. Chem. 2015, 58, 8616–8637. [Google Scholar] [CrossRef] [PubMed]
  245. Daniele, S.; Giacomelli, C.; Martini, C. Brain ageing and neurodegenerative disease: The role of cellular waste management. Biochem. Pharmacol. 2018, 158, 207–216. [Google Scholar] [CrossRef] [PubMed]
  246. Sands, W.A.; Page, M.M.; Selman, C. Proteostasis and ageing: Insights from long-lived mutant mice. J. Physiol. 2017, 595, 6383–6390. [Google Scholar] [CrossRef] [PubMed]
  247. He, L.Q.; Lu, J.H.; Yue, Z.Y. Autophagy in ageing and ageing-associated diseases. Acta Pharmacol. Sin. 2013, 34, 605–611. [Google Scholar] [CrossRef] [Green Version]
  248. Labbadia, J.; Morimoto, R.I. The biology of proteostasis in aging and disease. Annu. Rev. Biochem. 2015, 84, 435–464. [Google Scholar] [CrossRef]
  249. Jung, T.; Grune, T. Structure of the proteasome. Prog. Mol. Biol. Transl. Sci. 2012, 109, 1–39. [Google Scholar]
  250. Wilhelm, T.; Richly, H. Autophagy during ageing – from Dr Jekyll to Mr Hyde. FEBS J. 2018, 285, 2367–2376. [Google Scholar] [CrossRef]
  251. Li, W.W.; Li, J.; Bao, J.K. Microautophagy: Lesser-known self-eating. Cell. Mol. Life Sci. Cmls 2012, 69, 1125–1136. [Google Scholar] [CrossRef]
  252. Dice, J.F. Chaperone-Mediated Autophagy. Autophagy 2007, 3, 295–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Feng, Y.; He, D.; Yao, Z.; Klionsky, D.J. The machinery of macroautophagy. Cell Res. 2013, 24, 24. [Google Scholar] [CrossRef] [PubMed]
  254. Kamada, Y.; Sekito, T.; Ohsumi, Y. Autophagy in Yeast: ATOR-Mediated Response to Nutrient Starvation. In TOR: Target of Rapamycin; Thomas, G., Sabatini, D.M., Hall, M.N., Eds.; Springer: Berlin/Heidelberg, Germany, 2004; pp. 73–84. [Google Scholar]
  255. Settembre, C.; Ballabio, A. TFEB regulates autophagy: An integrated coordination of cellular degradation and recycling processes. Autophagy 2011, 7, 1379–1381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Sciarretta, S.; Yee, D.; Ammann, P.; Nagarajan, N.; Volpe, M.; Frati, G.; Sadoshima, J. Role of NADPH oxidase in the regulation of autophagy in cardiomyocytes. Clin. Sci. 2015, 128, 387–403. [Google Scholar] [CrossRef] [PubMed]
  257. Liu, T.; Ma, X.; Ouyang, T.; Chen, H.; Lin, J.; Liu, J.; Xiao, Y.; Yu, J.; Huang, Y. SIRT1 reverses senescence via enhancing autophagy and attenuates oxidative stress-induced apoptosis through promoting p53 degradation. Int. J. Biol. Macromol. 2018, 117, 225–234. [Google Scholar] [CrossRef] [PubMed]
  258. Zhou, J.; Liao, W.; Yang, J.; Ma, K.; Li, X.; Wang, Y.; Wang, D.; Wang, L.; Zhang, Y.; Yin, Y.; et al. FOXO3 induces FOXO1-dependent autophagy by activating the AKT1 signaling pathway. Autophagy 2012, 8, 1712–1723. [Google Scholar] [CrossRef] [Green Version]
  259. Höhn, A.; Jung, T.; Grimm, S.; Catalgol, B.; Weber, D.; Grune, T. Lipofuscin inhibits the proteasome by binding to surface motifs. Free Radic. Biol. Med. 2011, 50, 585–591. [Google Scholar] [CrossRef]
  260. Singh, R.; Kaushik, S.; Wang, Y.; Xiang, Y.; Novak, I.; Komatsu, M.; Tanaka, K.; Cuervo, A.M.; Czaja, M.J. Autophagy regulates lipid metabolism. Nature 2009, 458, 1131. [Google Scholar] [CrossRef]
  261. Kaushik, S.; Cuervo, A.M. Degradation of lipid droplet-associated proteins by chaperone-mediated autophagy facilitates lipolysis. Nat. Cell Biol. 2015, 17, 759. [Google Scholar] [CrossRef]
  262. Koseoglu, M.M.; Norambuena, A.; Sharlow, E.R.; Lazo, J.S.; Bloom, G.S. Aberrant Neuronal Cell Cycle Re-Entry: The Pathological Confluence of Alzheimer’s Disease and Brain Insulin Resistance, and Its Relation to Cancer. J. Alzheimer Dis. 2019, 67, 1–11. [Google Scholar] [CrossRef]
  263. Birdsall, V.; Waites, C.L. Autophagy at the synapse. Neurosci. Lett. 2019, 697, 24–28. [Google Scholar] [CrossRef] [PubMed]
  264. Hansen, M.; Rubinsztein, D.C.; Walker, D.W. Autophagy as a promoter of longevity: Insights from model organisms. Nat. Rev. Mol. Cell Biol. 2018, 19, 579–593. [Google Scholar] [CrossRef] [PubMed]
  265. Luna-Guevara, M.L.; Luna-Guevara, J.J.; Hernández-Carranza, P.; Ruíz-Espinosa, H.; Ochoa-Velasco, C.E. Chapter 3 - Phenolic Compounds: A Good Choice Against Chronic Degenerative Diseases. In Studies in Natural Products Chemistry; Attaur, R., Ed.; Elsevier: Amsterdam, The Netherlands, 2018; Volume 59, pp. 79–108. [Google Scholar]
  266. Manach, C.; Scalbert, A.; Morand, C.; Rémésy, C.; Jiménez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747. [Google Scholar] [CrossRef]
  267. Ferreira, I.C.; Barros, L.; Abreu, R. Antioxidants in wild mushrooms. Curr. Med. Chem. 2009, 16, 1543–1560. [Google Scholar] [CrossRef] [PubMed]
  268. Gil-Ramírez, A.; Pavo-Caballero, C.; Baeza, E.; Baenas, N.; Garcia-Viguera, C.; Marín, F.R.; Soler-Rivas, C. Mushrooms do not contain flavonoids. J. Funct. Foods 2016, 25, 1–13. [Google Scholar] [CrossRef] [Green Version]
  269. Surguchov, A.; Emamzadeh, F.N.; Surguchev, A.A. Amyloidosis and longevity: A lesson from plants. Biology 2019, 8, 43. [Google Scholar] [CrossRef]
  270. Sandoval-Acuña, C.; Ferreira, J.; Speisky, H. Polyphenols and mitochondria: An update on their increasingly emerging ROS-scavenging independent actions. Arch. Biochem. Biophys. 2014, 559, 75–90. [Google Scholar] [CrossRef]
  271. Kim, M.Y.; Seguin, P.; Ahn, J.K.; Kim, J.J.; Chun, S.C.; Kim, E.H.; Seo, S.H.; Kang, E.Y.; Kim, S.L.; Park, Y.J.; et al. Phenolic compound concentration and antioxidant activities of edible and medicinal mushrooms from Korea. J. Agric. Food Chem. 2008, 56, 7265–7270. [Google Scholar] [CrossRef]
  272. Heleno, S.A.; Barros, L.; Martins, A.; Queiroz, M.J.R.P.; Santos-Buelga, C.; Ferreira, I.C.F.R. Phenolic, polysaccharidic, and lipidic fractions of mushrooms from northeastern portugal: Chemical compounds with antioxidant properties. J. Agric. Food Chem. 2012, 60, 4634–4640. [Google Scholar] [CrossRef]
  273. Zhang, D.D.; Hannink, M. Distinct Cysteine Residues in Keap1 Are Required for Keap1-Dependent Ubiquitination of Nrf2 and for Stabilization of Nrf2 by Chemopreventive Agents and Oxidative Stress. Mol. Cell. Biol. 2003, 23, 8137. [Google Scholar] [CrossRef]
  274. Matzinger, M.; Fischhuber, K.; Heiss, E.H. Activation of Nrf2 signaling by natural products-can it alleviate diabetes? Biotechnol. Adv. 2018, 36, 1738–1767. [Google Scholar] [CrossRef] [PubMed]
  275. Martínez-Huélamo, M.; Rodríguez-Morató, J.; Boronat, A.; de la Torre, R. Modulation of Nrf2 by Olive Oil and Wine Polyphenols and Neuroprotection. Antioxid 2017, 6, 73. [Google Scholar] [CrossRef] [PubMed]
  276. Scapagnini, G.; Vasto, S.; Abraham, N.G.; Caruso, C.; Zella, D.; Fabio, G. Modulation of Nrf2/ARE pathway by food polyphenols: A nutritional neuroprotective strategy for cognitive and neurodegenerative disorders. Mol Neurobiol 2011, 44, 192–201. [Google Scholar] [CrossRef] [PubMed]
  277. Yahfoufi, N.; Alsadi, N.; Jambi, M.; Matar, C. The Immunomodulatory and Anti-Inflammatory Role of Polyphenols. Nutrients 2018, 10, 1618. [Google Scholar] [CrossRef] [PubMed]
  278. Nayernia, Z.; Jaquet, V.; Krause, K.-H. New insights on NOX enzymes in the central nervous system. Antioxid Redox Signal 2014, 20, 2815–2837. [Google Scholar] [CrossRef] [PubMed]
  279. Gandhi, S.; Abramov, A.Y. Mechanism of Oxidative Stress in Neurodegeneration. Oxidative Med. Cell. Longev. 2012, 2012, 11. [Google Scholar] [CrossRef] [PubMed]
  280. Cohen, G.; Kesler, N. Monoamine Oxidase and Mitochondrial Respiration. J. Neurochem. 1999, 73, 2310–2315. [Google Scholar] [CrossRef]
  281. Dos Santos, T.W.; Pereira, Q.C.; Teixeira, L.; Gambero, A.; Villena, J.A.; Ribeiro, M.L. Effects of polyphenols on thermogenesis and mitochondrial biogenesis. Int. J. Mol. Sci. 2018, 19, 5757. [Google Scholar] [CrossRef]
  282. Dong, W.; Wang, F.; Guo, W.; Zheng, X.; Chen, Y.; Zhang, W.; Shi, H. Aβ25–35 Suppresses Mitochondrial Biogenesis in Primary Hippocampal Neurons. Cell. Mol. Neurobiol. 2016, 36, 83–91. [Google Scholar] [CrossRef]
  283. Xu, W.; Barrientos, T.; Andrews, N.C. Iron and copper in mitochondrial diseases. Cell Metab. 2013, 17, 319–328. [Google Scholar] [CrossRef]
  284. Perez, C.A.; Wei, Y.; Guo, M. Iron-binding and anti-Fenton properties of baicalein and baicalin. J. Inorg. Biochem. 2009, 103, 326–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Refat, M.S. Synthesis and characterization of ligational behavior of curcumin drug towards some transition metal ions: Chelation effect on their thermal stability and biological activity. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2013, 105, 326–337. [Google Scholar] [CrossRef] [PubMed]
  286. Valdés, A.; Sullini, G.; Ibáñez, E.; Cifuentes, A.; García-Cañas, V. Rosemary polyphenols induce unfolded protein response and changes in cholesterol metabolism in colon cancer cells. J. Funct. Foods 2015, 15, 429–439. [Google Scholar] [CrossRef]
  287. Shen, M.; Chan, T.H.; Dou, Q.P. Targeting tumor ubiquitin-proteasome pathway with polyphenols for chemosensitization. Anti-Cancer Agents Med. Chem. 2012, 12, 891–901. [Google Scholar] [CrossRef]
  288. Pallauf, K.; Rimbach, G. Autophagy, polyphenols and healthy ageing. Ageing Res. Rev. 2013, 12, 237–252. [Google Scholar] [CrossRef] [PubMed]
  289. Rastogi, M.; Ojha, R.P.; Sagar, C.; Agrawal, A.; Dubey, G.P. Protective effect of curcuminoids on age-related mitochondrial impairment in female Wistar rat brain. Biogerontology 2014, 15, 21–31. [Google Scholar] [CrossRef]
  290. Chung, S.; Yao, H.; Caito, S.; Hwang, J.W.; Arunachalam, G.; Rahman, I. Regulation of SIRT1 in cellular functions: Role of polyphenols. Arch. Biochem. Biophys. 2010, 501, 79–90. [Google Scholar] [CrossRef] [Green Version]
  291. Horio, Y. Elucidation of the roles of protein deacetylase SIRT1 in health and diseases. Sapporo Med. J. 2018, 87, 1–8. [Google Scholar]
  292. Ren, Z.; He, H.; Zuo, Z.; Xu, Z.; Wei, Z.; Deng, J. The role of different SIRT1-mediated signaling pathways in toxic injury. Cell. Mol. Biol. Lett. 2019, 24, 36. [Google Scholar] [CrossRef]
  293. Maiese, K. The mechanistic target of rapamycin (mTOR) and the silent mating-type information regulation 2 homolog 1 (SIRT1): Oversight for neurodegenerative disorders. Biochem. Soc. Trans. 2018, 46, 351–360. [Google Scholar] [CrossRef]
  294. Bao, J.; Zheng, L.; Zhang, Q.; Li, X.; Zhang, X.; Li, Z.; Bai, X.; Zhang, Z.; Huo, W.; Zhao, X.; et al. Deacetylation of TFEB promotes fibrillar Aβ degradation by upregulating lysosomal biogenesis in microglia. Protein Cell 2016, 7, 417–433. [Google Scholar] [CrossRef] [PubMed]
  295. Palmieri, M.; Impey, S.; Kang, H.; di Ronza, A.; Pelz, C.; Sardiello, M.; Ballabio, A. Characterization of the CLEAR network reveals an integrated control of cellular clearance pathways. Hum. Mol. Genet. 2011, 20, 3852–3866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Brunet, A.; Sweeney, L.B.; Sturgill, J.F.; Chua, K.F.; Greer, P.L.; Lin, Y.; Tran, H.; Ross, S.E.; Mostoslavsky, R.; Cohen, H.Y.; et al. Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science 2004, 303, 2011–2015. [Google Scholar] [CrossRef] [PubMed]
  297. Pillarisetti, S. A Review of Sirt1 and Sirt1 Modulators in Cardiovascular and Metabolic Diseases. Recent Pat. Cardiovasc. Drug Discov. 2008, 3, 156–164. [Google Scholar] [CrossRef]
  298. Allard, J.S.; Perez, E.; Zou, S.; de Cabo, R. Dietary activators of Sirt1. Mol. Cell. Endocrinol. 2009, 299, 58–63. [Google Scholar] [CrossRef] [Green Version]
  299. Nardiello, P.; Pantano, D.; Lapucci, A.; Stefani, M.; Casamenti, F. Diet Supplementation with Hydroxytyrosol Ameliorates Brain Pathology and Restores Cognitive Functions in a Mouse Model of Amyloid-β Deposition. J. Alzheimer Dis. 2018, 63, 1161–1172. [Google Scholar] [CrossRef]
  300. Pantano, D.; Luccarini, I.; Nardiello, P.; Servili, M.; Stefani, M.; Casamenti, F. Oleuropein aglycone and polyphenols from olive mill waste water ameliorate cognitive deficits and neuropathology. Br. J. Clin. Pharmacol. 2017, 83, 54–62. [Google Scholar] [CrossRef]
  301. Cordero, J.G.; García-Escudero, R.; Avila, J.; Gargini, R.; García-Escudero, V. Benefit of Oleuropein Aglycone for Alzheimer’s Disease by Promoting Autophagy. Oxidative Med. Cell. Longev. 2018, 2018, 5010741. [Google Scholar] [CrossRef]
  302. Kim, H.S.; Quon, M.J.; Kim, J.A. New insights into the mechanisms of polyphenols beyond antioxidant properties; lessons from the green tea polyphenol, epigallocatechin 3-gallate. Redox Biol 2014, 2, 187–195. [Google Scholar] [CrossRef] [Green Version]
  303. Ulakcsai, Z.; Bagaméry, F.; Szökő, É.; Tábi, T. The role of autophagy induction in the mechanism of cytoprotective effect of resveratrol. Eur. J. Pharm. Sci. 2018, 123, 135–142. [Google Scholar] [CrossRef]
  304. Song, J.; Huang, Y.; Zheng, W.; Yan, J.; Cheng, M.; Zhao, R.; Chen, L.; Hu, C.; Jia, W. Resveratrol reduces intracellular reactive oxygen species levels by inducing autophagy through the AMPK-mTOR pathway. Front. Med. 2018, 12, 697–706. [Google Scholar] [CrossRef] [PubMed]
  305. Velagapudi, R.; Lepiarz, I.; El-Bakoush, A.; Katola, F.O.; Bhatia, H.; Fiebich, B.L.; Olajide, O.A. Induction of Autophagy and Activation of SIRT-1 Deacetylation Mechanisms Mediate Neuroprotection by the Pomegranate Metabolite Urolithin A in BV2 Microglia and Differentiated 3D Human Neural Progenitor Cells. Mol. Nutr. Food Res. 2019, 63, 1801237. [Google Scholar] [CrossRef] [PubMed]
  306. Sarubbo, F.; Ramis, M.R.; Kienzer, C.; Aparicio, S.; Esteban, S.; Miralles, A.; Moranta, D. Chronic Silymarin, Quercetin and Naringenin Treatments Increase Monoamines Synthesis and Hippocampal Sirt1 Levels Improving Cognition in Aged Rats. J. Neuroimmune Pharmacol. 2018, 13, 24–38. [Google Scholar] [CrossRef] [PubMed]
  307. Putteeraj, M.; Lim, W.L.; Teoh, S.L.; Yahaya, M.F. Flavonoids and its neuroprotective effects on brain ischemia and neurodegenerative diseases. Curr. Drug Targets 2018, 19, 1710–1720. [Google Scholar] [CrossRef] [PubMed]
  308. Huang, S.M.; Tsai, S.Y.; Lin, J.A.; Wu, C.H.; Yen, G.C. Cytoprotective effects of hesperetin and hesperidin against amyloid ss-induced impairment of glucose transport through downregulation of neuronal autophagy. Mol. Nutr. Food Res. 2012, 56, 601–609. [Google Scholar] [CrossRef] [PubMed]
  309. Xiao, J.B.; Högger, P. Dietary polyphenols and type 2 diabetes: Current insights and future perspectives. Curr. Med. Chem. 2015, 22, 23–38. [Google Scholar] [CrossRef]
  310. Rodríguez-García, C.; Sánchez-Quesada, C.; Gaforio, J.J.; Gaforio, J.J. Dietary flavonoids as cancer chemopreventive agents: An updated review of human studies. Antioxidants 2019, 8, 137. [Google Scholar] [CrossRef]
  311. Carrasco-Pozo, C.; Cires, M.J.; Gotteland, M. Quercetin and Epigallocatechin Gallate in the Prevention and Treatment of Obesity: From Molecular to Clinical Studies. J. Med. Food 2019, 22, 753–770. [Google Scholar] [CrossRef]
  312. Zang, M.; Xu, S.; Maitland-Toolan, K.A.; Zuccollo, A.; Hou, X.; Jiang, B.; Wierzbicki, M.; Verbeuren, T.J.; Cohen, R.A. Polyphenols Stimulate AMP-Activated Protein Kinase, Lower Lipids, and Inhibit Accelerated Atherosclerosis in Diabetic LDL Receptor–Deficient Mice. Diabetes 2006, 55, 2180. [Google Scholar] [CrossRef]
  313. Huang, J.; Zhang, Y.; Zhou, Y.; Zhang, Z.; Xie, Z.; Zhang, J.; Wan, X. Green Tea Polyphenols Alleviate Obesity in Broiler Chickens through the Regulation of Lipid-Metabolism-Related Genes and Transcription Factor Expression. J. Agric. Food Chem. 2013, 61, 8565–8572. [Google Scholar] [CrossRef]
  314. Bigagli, E.; Toti, S.; Lodovici, M.; Giovannelli, L.; Cinci, L.; D’Ambrosio, M.; Luceri, C. Dietary extra-virgin olive oil polyphenols do not attenuate colon inflammation in transgenic HLAB-27 rats but exert hypocholesterolemic effects through the modulation of HMGCR and PPAR-α gene expression in the liver. Lifestyle Genom. 2019, 11, 99–108. [Google Scholar] [CrossRef] [PubMed]
  315. Barquissau, V.; Ghandour, R.A.; Ailhaud, G.; Klingenspor, M.; Langin, D.; Amri, E.Z.; Pisani, D.F. Control of adipogenesis by oxylipins, GPCRs and PPARs. Biochimie 2017, 136, 3–11. [Google Scholar] [CrossRef] [PubMed]
  316. Wang, S.; Zhang, Q.; Zhang, Y.; Shen, C.; Wang, Z.; Wu, Q.; Zhang, Y.; Li, S.; Qiao, Y. Agrimol B suppresses adipogenesis through modulation of SIRT1-PPAR gamma signal pathway. Biochem. Biophys. Res. Commun. 2016, 477, 454–460. [Google Scholar] [CrossRef] [PubMed]
  317. Simons, K.; Ehehalt, R. Cholesterol, lipid rafts, and disease. J. Clin. Investig. 2002, 110, 597–603. [Google Scholar] [CrossRef] [PubMed]
  318. Salminen, A.; Kauppinen, A.; Suuronen, T.; Kaarniranta, K. SIRT1 longevity factor suppresses NF-κB-driven immune responses: Regulation of aging via NF-κB acetylation? BioEssays 2008, 30, 939–942. [Google Scholar] [CrossRef]
  319. Morgan, M.J.; Liu, Z.G. Crosstalk of reactive oxygen species and NF-κB signaling. Cell Res. 2011, 21, 103–115. [Google Scholar] [CrossRef]
  320. Scalbert, A.; Johnson, I.T.; Saltmarsh, M. Polyphenols: Antioxidants and beyond. Am. J. Clin. Nutr. 2005, 81, 215S–217S. [Google Scholar] [CrossRef]
  321. Rahman, I.; Chung, S. Dietary Polyphenols, Deacetylases and Chromatin Remodeling in Inflammation. World Rev. Nutr. Diet. 2010, 101, 84–94. [Google Scholar]
  322. Kostomoiri, M.; Fragkouli, A.; Sagnou, M.; Skaltsounis, L.A.; Pelecanou, M.; Tsilibary, E.C.; Τzinia, A.K. Oleuropein, an Anti-oxidant Polyphenol Constituent of Olive Promotes α-Secretase Cleavage of the Amyloid Precursor Protein (AβPP). Cell. Mol. Neurobiol. 2013, 33, 147–154. [Google Scholar] [CrossRef]
  323. Porzoor, A.; Alford, B.; Hügel, H.M.; Grando, D.; Caine, J.; Macreadie, I. Anti-amyloidogenic properties of some phenolic compounds. Biomolecules 2015, 5, 505–527. [Google Scholar] [CrossRef]
  324. Wobst, H.J.; Sharma, A.; Diamond, M.I.; Wanker, E.E.; Bieschke, J. The green tea polyphenol (-)-epigallocatechin gallate prevents the aggregation of tau protein into toxic oligomers at substoichiometric ratios. FEBS Lett. 2015, 589, 77–83. [Google Scholar] [CrossRef] [PubMed]
  325. Lee, D.H.; Lee, D.H.; Lee, J.S. Characterization of a new antidementia β-secretase inhibitory peptide from Saccharomyces cerevisiae. Enzym. Microb. Technol. 2007, 42, 83–88. [Google Scholar] [CrossRef]
  326. Park, I.H.; Jeon, S.Y.; Lee, H.J.; Kim, S.I.; Song, K.S. A β-secretase (BACE1) inhibitor hispidin from the mycelial cultures of Phellinus linteus. Planta Med. 2004, 70, 143–146. [Google Scholar] [PubMed]
  327. Bennett, L.; Sheean, P.; Zabaras, D.; Head, R. Heat-stable components of wood ear mushroom, Auricularia polytricha (higher basidiomycetes), inhibit in vitro activity of beta secretase (BACE1). Int. J. Med. Mushrooms 2013, 15, 233–249. [Google Scholar] [CrossRef]
  328. Song, S.H.; Choi, S.M.; Kim, J.E.; Sung, J.E.; Lee, H.A.; Choi, Y.H.; Bae, C.J.; Choi, Y.W.; Hwang, D.Y. α-Isocubebenol alleviates scopolamine-induced cognitive impairment by repressing acetylcholinesterase activity. Neurosci. Lett. 2017, 638, 121–128. [Google Scholar] [CrossRef]
  329. Sameem, B.; Saeedi, M.; Mahdavi, M.; Shafiee, A. A review on tacrine-based scaffolds as multi-target drugs (MTDLs) for Alzheimer’s disease. Eur. J. Med. Chem. 2017, 128, 332–345. [Google Scholar] [CrossRef]
  330. Orhan, I.; Kartal, M.; Tosun, F.; Şener, B. Screening of various phenolic acids and flavonoid derivatives for their anticholinesterase potential. Z. Fur Nat. Sect. C J. Biosci. 2007, 62, 829–832. [Google Scholar] [CrossRef]
  331. Dgachi, Y.; Sokolov, O.; Luzet, V.; Godyń, J.; Panek, D.; Bonet, A.; Martin, H.; Iriepa, I.; Moraleda, I.; García-Iriepa, C.; et al. Tetrahydropyranodiquinolin-8-amines as new, non-hepatotoxic, antioxidant, and acetylcholinesterase inhibitors for Alzheimer’s disease therapy. Eur. J. Med. Chem. 2017, 126, 576–589. [Google Scholar] [CrossRef]
  332. Kuppusamy, A.; Arumugam, M.; George, S. Combining in silico and in vitro approaches to evaluate the acetylcholinesterase inhibitory profile of some commercially available flavonoids in the management of Alzheimer’s disease. Int. J. Biol. Macromol. 2017, 95, 199–203. [Google Scholar] [CrossRef]
  333. Nair, A.B.; Jacob, S. A simple practice guide for dose conversion between animals and human. J. Basic Clin. Pharm. 2016, 7, 27–31. [Google Scholar] [CrossRef] [Green Version]
  334. Renaud, J.; Martinoli, M.-G. Considerations for the Use of Polyphenols as Therapies in Neurodegenerative Diseases. Int J Mol Sci 2019, 20, 1883. [Google Scholar] [CrossRef] [PubMed]
  335. Margină, D.; Ilie, M.; Grădinaru, D.; Androutsopoulos, V.P.; Kouretas, D.; Tsatsakis, A.M. Natural products—friends or foes? Toxicol. Lett. 2015, 236, 154–167. [Google Scholar] [CrossRef]
  336. Hodek, P.; Trefil, P.; Stiborová, M. Flavonoids-potent and versatile biologically active compounds interacting with cytochromes P450. Chem. Biol. Interact. 2002, 139, 1–21. [Google Scholar] [CrossRef]
  337. Bailey, D.G.; Malcolm, J.; Arnold, O.; Spence, J.D. Grapefruit juice–drug interactions. Br. J. Clin. Pharmacol. 1998, 46, 101–110. [Google Scholar] [CrossRef] [PubMed]
  338. Doostdar, H.; Burke, M.D.; Mayer, R.T. Bioflavonoids: Selective substrates and inhibitors for cytochrome P450 CYP1A and CYP1B1. Toxicology 2000, 144, 31–38. [Google Scholar] [CrossRef]
  339. Boocock, D.J.; Faust, G.E.; Patel, K.R.; Schinas, A.M.; Brown, V.A.; Ducharme, M.P.; Booth, T.D.; Crowell, J.A.; Perloff, M.; Gescher, A.J. Phase I dose escalation pharmacokinetic study in healthy volunteers of resveratrol, a potential cancer chemopreventive agent. Cancer Epidemiol. Prev. Biomark. 2007, 16, 1246–1252. [Google Scholar] [CrossRef]
  340. Vaz-da-Silva, M.; Loureiro, A.; Falcao, A.; Nunes, T.; Rocha, J.; Fernandes-Lopes, C.; Soares, E.; Wright, L.; Almeida, L.; Soares-da-Silva, P. Effect of food on the pharmacokinetic profile of trans-resveratrol. Int. J. Clin. Pharm. 2008, 46, 564–570. [Google Scholar] [CrossRef]
  341. Almeida, L.; Vaz-da-Silva, M.; Falcão, A.; Soares, E.; Costa, R.; Loureiro, A.I.; Fernandes-Lopes, C.; Rocha, J.F.; Nunes, T.; Wright, L. Pharmacokinetic and safety profile of trans-resveratrol in a rising multiple-dose study in healthy volunteers. Mol. Nutr. Food Res. 2009, 53, S7–S15. [Google Scholar] [CrossRef]
  342. Chow, H.S.; Cai, Y.; Hakim, I.A.; Crowell, J.A.; Shahi, F.; Brooks, C.A.; Dorr, R.T.; Hara, Y.; Alberts, D.S. Pharmacokinetics and safety of green tea polyphenols after multiple-dose administration of epigallocatechin gallate and polyphenon E in healthy individuals. Clin. Cancer Res. 2003, 9, 3312–3319. [Google Scholar]
  343. Herrschaft, H.; Nacu, A.; Likhachev, S.; Sholomov, I.; Hoerr, R.; Schlaefke, S. Ginkgo biloba extract EGb 761® in dementia with neuropsychiatric features: A randomised, placebo-controlled trial to confirm the efficacy and safety of a daily dose of 240 mg. J. Psychiatr. Res. 2012, 46, 716–723. [Google Scholar] [CrossRef]
  344. Ringman, J.M.; Frautschy, S.A.; Teng, E.; Begum, A.N.; Bardens, J.; Beigi, M.; Gylys, K.H.; Badmaev, V.; Heath, D.D.; Apostolova, L.G. Oral curcumin for Alzheimer’s disease: Tolerability and efficacy in a 24-week randomized, double blind, placebo-controlled study. Alzheimer Res. Ther. 2012, 4, 43. [Google Scholar] [CrossRef] [PubMed]
  345. Ihl, R.; Tribanek, M.; Bachinskaya, N.; Group, G.S. Efficacy and tolerability of a once daily formulation of Ginkgo biloba extract EGb 761® in Alzheimer’s disease and vascular dementia: Results from a randomised controlled trial. Pharmacopsychiatry 2012, 45, 41–46. [Google Scholar] [CrossRef] [PubMed]
  346. Figueira, I.; Menezes, R.; Macedo, D.; Costa, I.; Nunes dos Santos, C. Polyphenols beyond barriers: A glimpse into the brain. Curr. Neuropharmacol 2017, 15, 562–594. [Google Scholar] [CrossRef] [PubMed]
  347. Moussa, C.; Hebron, M.; Huang, X.; Ahn, J.; Rissman, R.A.; Aisen, P.S.; Turner, R.S. Resveratrol regulates neuro-inflammation and induces adaptive immunity in Alzheimer’s disease. J. Neuroinflamm. 2017, 14, 1. [Google Scholar] [CrossRef] [PubMed]
  348. Turner, R.S.; Thomas, R.G.; Craft, S.; van Dyck, C.H.; Mintzer, J.; Reynolds, B.A.; Brewer, J.B.; Rissman, R.A.; Raman, R.; Aisen, P.S. A randomized, double-blind, placebo-controlled trial of resveratrol for Alzheimer disease. Neurology 2015, 85, 1383–1391. [Google Scholar] [CrossRef] [PubMed]
  349. Beck, S.M.; Ruge, H.; Schindler, C.; Burkart, M.; Miller, R.; Kirschbaum, C.; Goschke, T. Effects of Ginkgo biloba extract EGb 761® on cognitive control functions, mental activity of the prefrontal cortex and stress reactivity in elderly adults with subjective memory impairment–a randomized double-blind placebo-controlled trial. Hum. Psychopharmacol. Clin. Exp. 2016, 31, 227–242. [Google Scholar] [CrossRef] [PubMed]
  350. Levin, J.; Maaß, S.; Schuberth, M.; Respondek, G.; Paul, F.; Mansmann, U.; Oertel, W.H.; Lorenzl, S.; Krismer, F.; Seppi, K. The PROMESA-protocol: Progression rate of multiple system atrophy under EGCG supplementation as anti-aggregation-approach. J. Neural Transm. 2016, 123, 439–445. [Google Scholar] [CrossRef]
  351. Pandareesh, M.; Mythri, R.; Bharath, M.S. Bioavailability of dietary polyphenols: Factors contributing to their clinical application in CNS diseases. Neurochem. Int. 2015, 89, 198–208. [Google Scholar] [CrossRef]
  352. Hu, M.; Wu, B.; Liu, Z. Bioavailability of polyphenols and flavonoids in the era of precision medicine. Mol. Pharm. 2017, 14, 2861–2863. [Google Scholar] [CrossRef]
  353. D Archivio, M.; Filesi, C.; Di Benedetto, R.; Gargiulo, R.; Giovannini, C.; Masella, R. Polyphenols, dietary sources and bioavailability. Ann. Ist. Super. Sanita 2007, 43, 348. [Google Scholar]
  354. Bravo, L. Polyphenols: Chemistry, dietary sources, metabolism, and nutritional significance. Nutr. Rev. 1998, 56, 317–333. [Google Scholar] [CrossRef] [PubMed]
  355. Scalbert, A.; Morand, C.; Manach, C.; Rémésy, C. Absorption and metabolism of polyphenols in the gut and impact on health. Biomed. Pharmacother. 2002, 56, 276–282. [Google Scholar] [CrossRef]
  356. Cermak, R.; Wolffram, S. The potential of flavonoids to influence drug metabolism and pharmacokinetics by local gastrointestinal mechanisms. Curr. Drug Metab. 2006, 7, 729–744. [Google Scholar] [CrossRef]
  357. Ballabh, P.; Braun, A.; Nedergaard, M. The blood–brain barrier: An overview: Structure, regulation, and clinical implications. Neurobiol. Dis. 2004, 16, 1–13. [Google Scholar] [CrossRef] [PubMed]
  358. Figueira, I.; Garcia, G.; Pimpão, R.C.; Terrasso, A.; Costa, I.; Almeida, A.; Tavares, L.; Pais, T.; Pinto, P.; Ventura, M. Polyphenols journey through blood-brain barrier towards neuronal protection. Sci. Rep. 2017, 7, 11456. [Google Scholar] [CrossRef] [PubMed]
  359. Squillaro, T.; Cimini, A.; Peluso, G.; Giordano, A.; Melone, M.A.B. Nano-delivery systems for encapsulation of dietary polyphenols: An experimental approach for neurodegenerative diseases and brain tumors. Biochem. Pharmacol. 2018, 154, 303–317. [Google Scholar] [CrossRef] [PubMed]
  360. Prado-Audelo, D.; María, L.; Caballero-Florán, I.H.; Meza-Toledo, J.A.; Mendoza-Muñoz, N.; González-Torres, M.; Florán, B.; Cortés, H.; Leyva-Gómez, G. Formulations of curcumin nanoparticles for brain diseases. Biomolecules 2019, 9, 56. [Google Scholar] [CrossRef]
  361. Yavarpour-Bali, H.; Ghasemi-Kasman, M.; Pirzadeh, M. Curcumin-loaded nanoparticles: A novel therapeutic strategy in treatment of central nervous system disorders. Int. J. Nanomed. 2019, 14, 4449. [Google Scholar] [CrossRef]
  362. Davidov-Pardo, G.; McClements, D.J. Resveratrol encapsulation: Designing delivery systems to overcome solubility, stability and bioavailability issues. Trends Food Sci. Technol. 2014, 38, 88–103. [Google Scholar] [CrossRef]
Figure 1. Current drug targets and molecular events occurring in the Alzheimer’s Disease (AD) brain microenvironment.
Figure 1. Current drug targets and molecular events occurring in the Alzheimer’s Disease (AD) brain microenvironment.
Ijms 20 05090 g001
Figure 2. Structures of some polyphenols that show neuroprotective functions against AD.
Figure 2. Structures of some polyphenols that show neuroprotective functions against AD.
Ijms 20 05090 g002
Figure 3. Schematic representation for showing molecular mechanisms by which polyphenols acts as antioxidants (adapted from [36,270,273,274,277,281,284]).
Figure 3. Schematic representation for showing molecular mechanisms by which polyphenols acts as antioxidants (adapted from [36,270,273,274,277,281,284]).
Ijms 20 05090 g003
Figure 4. SIRT1 activation by polyphenols and its effect in protein degradation pathways in the intracellular environment (adapted from [37,290,291,292,293,296]).
Figure 4. SIRT1 activation by polyphenols and its effect in protein degradation pathways in the intracellular environment (adapted from [37,290,291,292,293,296]).
Ijms 20 05090 g004
Figure 5. Modulation of longevity by the action of polyphenols through SIRT1 activation (adapted from [288,291,292,295,296,297,298]).
Figure 5. Modulation of longevity by the action of polyphenols through SIRT1 activation (adapted from [288,291,292,295,296,297,298]).
Ijms 20 05090 g005
Table 1. Neuroprotective roles of some polyphenols for AD.
Table 1. Neuroprotective roles of some polyphenols for AD.
PolyphenolAnalytical System EPCa/ROAbEffects of Polyphenols at Cellular LevelEffects in Relation to ADReference
QuercetinIn vitroNAmTORC inhibitorInduces autophagy, anti-amyloidogenic, inhibits proteasomal degradation, antioxidant, restores biometal distribution, antiproliferative and enhances neuronal synapsis[4,5,6,7,8]
ARPE 19 cells2 μMTFEB activation
APPswe cells10 μMInhibits Aβ fibril formation
Rat neonatal cardiomyocytes5 μMInhibits all the catalytic subunits of proteasome
In vitroNAChelates iron
In vitroNAReduces ROS and RNS
In silico and in vitroNAInhibits acetyl choline esterase
ResveratrolTg6799 mice60 mg/kg/d for 60 d/oral administrationReduces amyloid plaque formationInduces autophagy, increases lysosomal biogenesis, restores lipid homeostasis, increases stress resistance, regulates cell cycle, antiproliferative, anti-apoptotic, increases longevity and anti-inflammatory[9,10,11,12,13,14]
Primary neuronal culture30 μMSIRT1 activation and NFκB inhibition
Obese healthy men clinical trial150 mg/d for 30 d/oral administrationTFEB activation
Human aortic endothelial cells50 μMAMPK mediated LC3II activation
Human aortic endothelial cells10 μMDecreases ROS and RNS, increases SOD
LNCaP cells20 μMp53 regulation, PI3K/Akt/mTOR inhibition, induces FOXO transcriptional activity including cell cycle regulation and stress resistance
Epigallocatechin gallate (EGCG)Human bladder cancer cell line T24 20 μg/mlInhibits Beclin1 suppressors and PI3K/Akt/mTORInduces autophagy, restores lipid homeostasis, anti-amyloidogenic, increases antioxidant capacity, restores impaired autophagosomes and biometal distribution, increases cell survival[15,16,17,18,19]
Bovine aortic endothelial cells10 μMIncreases LC3II formation and activates AMPK/ULK1
HepG2 cells40 μMDegrades lipid droplets through Ca2+/CAMKKB AMPK dependent mechanism
In vitroNAChelates zinc and copper
PC12 cells (rat pheochromocytoma100 μg/mLInteracts with Aβ40 and changes its conformation, inhibits lipofuscin formation
AnthocyaninSprague–Dawley rats100 mg/kg/d for 28 d/oral administrationRestores calcium homeostasis and activates Nrf2 subsequently activating phase II detoxifying genesActivates autophagy, increases expression of anti-oxidant genes, reduces ROS and increases cell survival[20,21,22,23]
HT22 cells and primary cultures of hippocampal neurons0.1 mg/mLInduces AMPK
In vitro0.005 mg/mLROS scavenging
HCC cell lines PLC/PRF/5 and HepG2 cells0.2 mg/mLIncrease expression of Beclin1, LC3 II
KaempferolSK-HEP-1 human hepatic cancer cell75 μMIncreases the levels of p-AMPK, LC3-II, Atg 5, Atg 7, Atg 12 and beclin 1, inhibits PI3K/Akt/mTORReduces mitochondrial dysfunction, anti-proliferative, increases autophagy, increases unfolded protein response, reduces APOE4 fragmentation and associated toxicity[24,25,26,27]
BALB/c nude mice150 mg/kg/d for 31 d/intraperitoneal injectionActivates DNMT methyltransferase ubiquitination
SCC-4, human tongue squamous cell carcinoma cell50 µMActivates IRE1-JNK-CHOP signaling, downregulates ERK1/2 signaling which reduces MMP2
HydroxytyrosolMale db/db (C57BL/6J) mice10 mg/kg/d for 8 weeks/oral administrationActivates Nrf2 and SIRT1/AMPK/PGC-1, reduces protein oxidation, increases NMDAR1 and NGF mRNA expressionEnhances autophagy, increases stress resistance and longevity, antioxidant, anti-inflammatory, restores lipid homeostasis and improves cognition[28,29,30,31,32,33]
VECs cells50 μMActivates AMPK/FOXO3a
VECs cells10 μMReduces ROS
VAFs from Sprague–Dawley rats25 μMIncreases LC3II/LC3I, Bcl1 and SIRT1 expression
HepG2 and Huh7 cells100 μMInhibits PI3K/Akt/mTOR, expression of IL1β & IL6, and NFκB DNA binding
Rat hepatocytes25 μMInhibits Acetyl CoA carboxylase, HMG CoA reductase, diacylglycerol acyl transferase
Oleuropein aglyconeRat ventricular myocyte100 μMIncreases Bcl1 and LC3II expression, TFEB nuclear localization, LAMP1 and p62 expressionInduces autophagy, increases lysosomal biogenesis and reduces oxidative damage[33,34,35]
Human SH-SY5Y neuroblastoma cells and rat RIN5F insulinoma cells50 μMInhibits MAOA, induces AMPK/ULK1, inhibits mTOR
Rat hepatocytes25 μMInhibits acetyl CoA carboxylase, HMG CoA reductase and diacylglycerol acyl transferase
CurcuminMale Sprague–Dawley rats15 mg/kg/d for 4 weeks/subcutaneous injectionActivates AMPK and regulates lipid metabolismInduces autophagy, restores lipid homeostasis, antioxidant, anti-amyloidogenic, anti-inflammatory, anti-apoptotic
antiproliferative, increases lysosomal biogenesis and longevity
[36,37,38,39,40,41,42,43,44,45]
Adult male Wistar rats30 mg/kg for 30 d/oral administrationActivates Nrf2, inhibits NFκB and mTOR
Adult Swiss male albino mice80 mg/kg/d for 7 d/intraperitoneal injectionInhibits MaoB and reduces ROS
APPswe Tg2576 transgenic mice (chronic 500 ppm curcumin diet)Blood curcumin level ~2 μM for 1 h/injection in right carotid arteryInhibits formation of Aβ, oligomers, fibrils and plaques
Tsc2+/+, Tsc2−/− MEFs and HCT116 cells10 μMActivates TFEB, increases levels of LC3 and inhibits pAkt
Sprague–Dawley rats’ primary cortical neurons10 μMUpregulates SIRT1 and inhibits Bax
APP/PS1 double transgenic mice160 ppm for 6 months/oral administrationInhibits PI3K/Akt/mTOR signaling, increases LC3I/II and Beclin1 expression
MyricetinHepG2 Cells50 μMInhibits mTOR and increases LC3II expressionInduces autophagy, antiproliferative, increases stress resistance, longevity, antioxidant capacity and mitochondrial regeneration[46,47,48]
Adipocytes differentiated from C3H10T1/2 cells10 μMActivates SIRT1/SIRT3/SIRT5
Male ICR mice50 mg/kg/d for 21 d/oral administrationIncreases mitochondrial mass and increases PGC1α, SIRT1, TFAM, Nrf1 & FOXO1
Urolithin AC2C12 myoblasts50 μMInduces mitophagy, increases LC3I/LC3II and activates AMPK signalingIncreases mitophagy, and autophagy, antioxidant, increases lysosomal biogenesis, anti-inflammatory, anti-amyloidogenic, improves cognition and longevity[49,50]
Female APP/PS1 transgenic mice B6C3-Tg (APPswe, PS1dE9) 85Dbo/J and age-matched wild type mice300 mg/kg/d for 14 d/oral administrationActivates AMPK, decreases NFκB/MAPK/BACE1 activities and APP levels
Ferulic AcidHeLa cells and mouse primary hepatocytes1 mMIncreases LC3 II and inhibits mTORAnti-apoptotic, anti-amyloidogenic, antioxidant, anti-inflammatory and induces autophagy[51,52,53,54]
In vitroNAInhibits Aβ aggregation and reduces ROS
(APP)swe/presenilin 1(PS1)dE9 (APP/PS1) mouse model5.3 mg/kg/d for 6 months/oral administrationReduces amyloid deposition and interleukin-1 beta (IL-1β) levels
AcacetinDrosophila melanogaster100 μMInhibits BACE1Anti-amyloidogenic, antioxidant, anti-inflammatory and induces autophagy[55,56,57,58]
C57BL/6J mice∼10 mg/kg/d for 14 d/oral administration (gavage)Inhibits MAPK and PI3K/Akt pathways
ICR mice100 mg/kg for 7 h/intraperitoneal injectionIncreases LC3II, Atg5 and Atg7 expression, modulates TNF-α/IL-6 expression and suppresses TLR4 signaling
BaicaleinSH-SY5Y human neuroblastoma cells12.5 μMIncreases ROS scavenging and activates Nrf2 Anti-amyloidogenic, anti-apoptotic, antioxidant, anti-inflammatory, inhibits excitotoxicity, stimulates neurogenesis and neuronal differentiation[59,60,61,62,63,64,65]
In vitroNAChelates iron
CHO/APPwt cells5 μMInduces α-secretase and inhibits Aβ formation
In vitro30 μMDissociates amyloid aggregates, Aβ oligomerization and fibrillation
HeLa cells100 μMInhibits NFκB activation
C57BL/6J APP/PS1 mice80 mg/kg/d for 60 d/oral administration (drinking water)Inhibits GSK3β mediated tau phosphorylation
Sprague-Dawley male rats20 mg/kg 30 min before and 2/4 h after onset of ischemia/intraperitoneal injectionInduces Bcl-2/Bcl-xL associated phosphorylation
IcariinPrimary cortical neurons prepared from E16-17 mouse embryos1.2 μMActivates SIRT1Antioxidant, anti-amyloidogenic, reduces ER stress, increases synapsis and neuronal plasticity, inhibits tau hyperphosphorylation, increases cell viability, antiapoptotic and anti-inflammatory[66,67,68,69,70,71,72,73]
Wistar rats60 mg/kg/d for 3 months/oral administrationIncreases SOD activity
Tg2576 mouse model60 mg/kg/d for 3 months/oral administrationReduces expression of BACE1 and APP
Sprague-Dawley rats120 mg/kg/d for 28 d/oral administrationInduces PSD95, BDNF, pTrkB, pAkt, and pCREB expression
SH-SY5Y cells1 μMInhibits GSK3β activation
PC12 cells10 μMInhibits JNK/p38, MAPK and p53 activity
HT29 and HCT11620 μMInhibits NFκB signaling
NobiletinMale 3XTg-AD mice30 mg/kg/d for 3 months/intraperitoneal injectionReduces Aβ levels and plaque formation in brainAnti-amyloidogenic, increases stress resistance, neuronal synapsis and plasticity, antioxidant and anti-inflammatory[74,75,76,77]
Male Sprague-Dawley rats25 mg/kg/d for 3d/intraperitoneal injectionIncreases activity of Akt, CREB, BDNF and Bcl2, increases Nrf2, HO-1, SOD1 and GSH expression, reduces NFκB, MMP-9 and MDA expression
GenisteinIn silico and in vitroNAInhibits chymotrypsin-like activity of proteasomesAntioxidant, increases degradation of Aβ, increases apoptosis, enhances autophagy and inhibits proteasomal protein degradation[78,79,80,81]
LNCaP cells100 μMIncreases Kip1 and reduces IκBα/Bax
Human dermal fibroblasts (HDFa)30 μMIncreases TFEB expression
Human mammary gland tumor cells (MCF-7)0.5 μMEnhances antioxidant gene expression
LuteolinHT-29 cells50 μMReduces ROS, NFκB signaling, Cox2 expression, blocks JAK/STAT signalingAnti-inflammatory, antioxidant, modulates autophagy and apoptosis, increases survival[82,83,84,85,86]
Male Sprague-Dawley rat myocytes8 μMDownregulates Bax expression, upregulates PI3k/Akt signaling and Bcl-2 expression
Human HCC cell line SMMC-772100 μMIncreases expression of LC3B-II, Bcl1 and caspase 8
MangiferinSwiss albino male rats15 mg/kg/d for 14 d/intraperitoneal injectionIncreases ROS scavenging, activates Nrf2, inhibits NFκB signaling, increases GSH levels, decreases lipid peroxidationAntioxidant, anti-apoptotic, chelates metals, increases stress resistance, autophagy, longevity, neuronal synapsis and plasticity[87,88,89,90,91]
In vitroNARescues mitochondrial respiration, chelates iron
Male Swiss albino mice40 mg/kg/d for 21 d/oral administrationReduces lipid peroxides and ROS/RNS induced by aluminum and restores regulation of BDNF and NGF
Human astroglioma U87MG, U373MG and CRT-MG cells100 μMInhibits PI3K/Akt signaling, MAPK pathway, MMP9 gene expression
Footnotes: NA, not applicable; a EPC, minimum concentration of the polyphenols that have significant neuroprotective effect; b ROA, route of administration of polyphenols in in vivo models.

Share and Cite

MDPI and ACS Style

Dhakal, S.; Kushairi, N.; Phan, C.W.; Adhikari, B.; Sabaratnam, V.; Macreadie, I. Dietary Polyphenols: A Multifactorial Strategy to Target Alzheimer’s Disease. Int. J. Mol. Sci. 2019, 20, 5090. https://doi.org/10.3390/ijms20205090

AMA Style

Dhakal S, Kushairi N, Phan CW, Adhikari B, Sabaratnam V, Macreadie I. Dietary Polyphenols: A Multifactorial Strategy to Target Alzheimer’s Disease. International Journal of Molecular Sciences. 2019; 20(20):5090. https://doi.org/10.3390/ijms20205090

Chicago/Turabian Style

Dhakal, Sudip, Naufal Kushairi, Chia Wei Phan, Benu Adhikari, Vikineswary Sabaratnam, and Ian Macreadie. 2019. "Dietary Polyphenols: A Multifactorial Strategy to Target Alzheimer’s Disease" International Journal of Molecular Sciences 20, no. 20: 5090. https://doi.org/10.3390/ijms20205090

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop